Ritual Violence in the Ancient Andes: Reconstructing Sacrifice on the North Coast of Peru 9781477310571

Traditions of sacrifice exist in almost every human culture and often embody a society’s most meaningful religious and s

236 43 111MB

English Pages 468 [472] Year 2016

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Ritual Violence in the Ancient Andes: Reconstructing Sacrifice on the North Coast of Peru
 9781477310571

Citation preview

Recto Runninghead  i

Ritual Vi o le n c e i n t h e A nc i en t An des

i i   Verso Runninghead

Recto Runninghead  i i i

Ritual Violence in the Ancient Andes

Reconstructing Sacrifice on the North Coast of Peru

E d i t e d by H a a g e n D . K l a u s a n d J . M a r l a T oyn e

university of texas press, austin

iv  Verso Runninghead

Copyright © 2016 by the University of Texas Press All rights reserved Printed in the United States of America First edition, 2016 Requests for permission to reproduce material from this work should be sent to: Permissions University of Texas Press P.O. Box 7819 Austin, TX 78713-7819 http://utpress.utexas.edu/index.php/rp-form ∞

The paper used in this book meets the minimum requirements of ANSI/NISO Z39.48-1992 (R1997) (Permanence of Paper).

Library of Congress Cataloging-in-Publication Data Names: Klaus, Haagen D., editor. Toyne, J. Marla, editor. Title: Ritual violence in the ancient Andes : reconstructing sacrifice on the north coast of Peru / edited by Haagen D. Klaus and J. Marla Toyne. Description: First edition. Austin : University of Texas Press, 2016. Includes bibliographical references and index. Series: The William and Bettye Nowlin series in art, history, and culture of the Western Hemisphere Identifiers: LCCN 2015039599   ISBN 9781477309377 (cloth : alk. paper)   ISBN 9781477309636 (pbk. : alk. paper)   ISBN 9781477310571 (library e-book)   ISBN 9781477310588 (non-library e-book) Subjects: LCSH: Indians of South America—Peru—Rites and ceremonies. Indians of South America—Peru—Antiquities. Sacrifice—Peru. Human sacrifice—Peru. Peru—Antiquities. Classification: LCC F3429.3.R58 R585 2016   DDC 985/.01—dc23   LC record available at http://lccn.loc.gov/2015039599 doi: 10.7560/309377

Recto Runninghead  v

To Gordon C. Pollard, who first put a trowel in my hand HDK ———

To Andrew Nelson, for demonstrating to me WHAT was SO important JMT ———

In memory of Christine Nelson JMT and HDK

THIS PAGE INTENTIONALLY LEFT BLANK

Recto Runninghead  vi i

con te n ts

Foreword  Clark Spencer Larsen  xi Preface and Acknowledgments xv chapter one  1

Ritual Violence on the North Coast of Peru: Perspectives and Prospects in the Archaeology of Ancient Andean Sacrifice Haagen D. Klaus and J. Marla Toyne

part one  25 Ancient Ritual Variation and Methodological Advances in Studies of Sacrifice Haagen D. Klaus and J. Marla Toyne chapter two  29 Ritual Killing, Mutilation, and Dismemberment at Huaca de la Luna: Sharp Force Trauma Among Moche Sacrifice Victims in Plazas 3A and 3C Laurel S. Hamilton chapter three  64 The Taphonomy of Ritual Killing on the North Coast of Peru: Perspectives from Huaca de la Luna and Pacatnamú Heather C. Backo chapter four  97 Ritual Strangulation in the Southern Moche World: Mortuary Ligatures as Tools of Liturgical Violation David Chicoine chapter five  120 Bodies and Blood: Middle Sicán Human Sacrifice in the Lambayeque Valley Complex (AD 900–1100) Haagen D. Klaus and Izumi Shimada

vi i i   Contents vi i i   Verso Runninghead

chapter six  150 Precious Gifts: Mortuary Patterns and the Shift from Animal to Human Sacrifice at Santa Rita B in the Middle Chao Valley, Peru Catherine Gaither, Jonathan Kent, Jonathan Bethard, Victor Vasquez, and Teresa Rosales chapter seven  178 Human Sacrifice at the Chotuna-Chornancap Archaeological Complex: Traditions and Transformations of Ritual Violence Under Chimú and Inka Rule Haagen D. Klaus, Bethany L. Turner, Fausto Saldaña, Samuel Castillo, and Carlos Wester

part two  211 Ancient Identities, Ambiguous Deaths, and Complex Burials Haagen D. Klaus and J. Marla Toyne chapter eight  215 Life Before Death: A Paleopathological Examination of Human Sacrifice at the Templo de la Piedra Sagrada, Túcume, Peru J. Marla Toyne chapter nine  244 The Killing of Captives on the North Coast of Peru in Pre-Hispanic Times: Iconographic and Bioarchaeological Evidence John W. Verano and Sara S. Phillips chapter ten  266 Reconsidering Retainers: Identity, Death, and Sacrifice in High-Status Funerary Contexts on the North Coast of Peru Sylvia Bentley and Haagen D. Klaus chapter eleven  291

Human Sacrifice: A View from San José de Moro Elsa Tomasto-cagigao, Mellisa Lund, Luis Jaime Castillo, and Lars Fehren-Schmitz

Recto Runninghead  Contents  i x i x

part three  315 Continuums of Killing: Sacrifice of Animals and Objects Haagen D. Klaus and J. Marla Toyne chapter twelve  319 Life Histories of Sacrificed Camelids from Huancaco (Virú Valley) Paul Szpak, Jean-François Millaire, Christine White, Steve Bourget, and Fred Longstaffe chapter thirteen  342 Posts and Pots: Propitiatory Ritual at Huaca Santa Clara in the Virú Valley, Peru Jean-François Millaire

part four  359 Perspectives from Beyond the North Coast of Peru Haagen D. Klaus and J. Marla Toyne chapter fourteen  361 Practicing and Performing Sacrifice Tiffiny a. Tung chapter fifteen  371 Mesoamerican Perspectives on the (Bio)archaeology of Andean Ritual Violence Vera Tiesler

Reference List 381 Index 451

THIS PAGE INTENTIONALLY LEFT BLANK

Recto Runninghead  xi

foreword

Human sacrifice has held the attention of the public for centuries. Its existence is detailed in folklore around the world, and it has been documented historically and archaeologically in a range of past societies. In teaching bioarchaeology, most students are horrified, yet fascinated, by the topic. The class discussion always gravitates to the same question: what motivates a society to allow ritual killing, sometimes involving those unable to defend themselves, such as children? This volume on ritual killing on the north coast of Peru provides arguably the most comprehensive dataset for addressing this fundamental question. The chapters collectively demonstrate that ritual killing is at once complex and variable. Deeper understandings of this ancient behavior may be found in the rich intersections between archaeological, bioarchaeological, historical, temporal, and spatial contexts. Importantly, ritual killing is not a bizarre, superstitious behavior exhibited by so-called primitive people. Rather, it is a profoundly revealing element of a society’s perception and interaction with the natural, cultural, and spiritual worlds. Much of the early literature on ritual killing in Andean South America has been dominated by an art history paradigm and case study orientations, both of which provided valuable, productive insights on sacrifice and descriptions of the osteological manifestations of trauma. This volume takes us from the realm of description to the contextualized reconstruction of once-living people. As with the world today, many of these societies shared common beliefs, some of which lasted many generations. The north coastal region of Andean South America is a remarkable laboratory for the study of human social behavior, largely owing to the robust archaeological and ethnohistoric records of death and ritual. Using a range of current and developing methods, the authors of the 15 chapters provide a succinct presentation of the development of sacrifice in the region over the last 2,000 years of history. They succeed in doing so by their collective development of new approaches to understanding the diversity and meaning of sacrifice. Collectively, they tell us that the motivations for ritual killing emerge from a complex web of social rules, beliefs, histories, politics, and identities. While the contextual bioarchaeological approach to human remains has emerged to complement the

x i i   Foreword xi i   Verso Runninghead

earlier approaches, this volume also takes some initial steps to further expand the concepts and definitions of Andean sacrifice—especially in how to better understand the ways in which animals and objects were also offered to the supernatural realm. Although the roots of ritual killing may extend further back into Andean prehistory, it is not until the rise of Moche culture in the early first millennium AD that we begin to see the extraordinary physical record of ritual violence begin to unfold, replete with evidence of sacrifice depicted in iconography. Various Moche contexts show highly patterned sacrifices involving throat-slicing and cranial trauma, and this skeletal record holds strong parallels to their iconography—but bioarchaeological studies of Moche sacrifice victims simultaneously record a previously undefined degree of ritual diversity just under the surface of previous understandings. Contrary to some earlier notions suggesting there was a lull in ritual violence following the Moche, the archaeological record now reveals the continuous presence of sacrifice in the later north coast cultures, especially the Sicán and Chimú. In at least one setting, the pattern of body treatment is largely the same, but it involved new innovations relating to age composition (increased presence of children) and highly formalized patterns of post-death treatment and placement of the victim’s remains. The period immediately preceding the arrival of Europeans in the sixteenth century also saw continued practice of ritual killing, blending long-held, conservative ritual patterns with new variations on the theme of sacrifice during consecutive eras of foreign imperial rule. This book establishes that the diversity and scale of sacrifice and treatment of the victims during these times was well beyond anything previously imagined. This volume succeeds in all of its goals. The authors collectively focus on context combined with innovative and creative combinations of methods. The patterns of violence so clearly displayed in ancient skeletons will pique the interest of most readers. Yet that is only a small part of the story of the history of ritual killing in this remarkable region of the world. This book makes clear that the description of the skeletal record of ritual killing and sacrifice is a relatively straightforward process, especially in regard to documenting key demographic attributes of age and sex and describing the location and type of trauma. I say this not intending to underplay this part of the record. In fact, the authors and editors go to great lengths to describe and illustrate these forms of evidence

x i i i xi ii RectoForeword  Runninghead 

using a notable and appropriate variety of sophisticated anthropological, archaeological, and biological tools. The more difficult challenge involves the extraction of the full contextual record in order to provide a broader understanding of the values of a society—and its motivations for behaviors that result in the intentional, and often violent, ritualized death of human beings. The book also makes clear that the anthropology of death and ritual, for this and other settings, is best studied not from just one traditional subdisciplinary corner of anthropology—whether biological anthropology, archaeology, or cultural anthropology—with each studying an isolated piece of the record. Each of these pieces alone does not provide the kind of compelling story that this book succeeds in telling. Indeed, this volume demonstrates the essential nature of intraand interdisciplinary collaboration in reconstructing the social and behavioral worlds of past people. Clark Spencer Larsen Department of Anthropology, The Ohio State University

THIS PAGE INTENTIONALLY LEFT BLANK

Recto Runninghead  xv

pr e face an d ackn owle dgments

The archaeological study of ritual violence on the north coast of Peru has burgeoned since the end of the twentieth century. Historically, the study of ritual killing had long been in the domain of art history and, to a lesser extent, ethnohistorians. Since the 1990s, discoveries of multiple settings of sacrifice fundamentally redefined the scope, perspective, methods, and questions surrounding this theme in Andean archaeology. At the same time, significant developments in bioarchaeological science also unfolded. This allowed for those investigating the remains of sacrifice victims, for the first time, to provide enhanced and nuanced analyses of ritual killing in the ancient Andes. In 2010, we decided to organize a symposium titled “Human Sacrifice on the North Coast of Peru: New Perspectives from Diachronic and Multidisciplinary Studies of Ancient Ritual Killing” to be held at the 76th annual meeting of the Society for American Archaeology in Sacramento, California, from March 30 to April 3, 2011. Our main objectives were to bring together the community of scholars researching and writing about sacrifice throughout the north coast of Peru. The majority of invitees were able to attend. By the end of the symposium, the speakers had constructively bridged previous generations of research styles with contemporary, empirical, and often bioarchaeologically based approaches. The participants demonstrated newly recognized temporal patterns and previously undefined variability of sacrifice across space in multiple north coast cultures. The papers presented at the symposium also broadened our perceptions of sacrifice beyond the act of killing. Some of this emerged from the application of new theoretical directions in mortuary archaeology. In addition, creative configurations of new analytical methods and regional or holistic research styles helped to define a range of pre-, peri-, and postsacrificial ritual processes. Another very important thread illustrated the conceptual expansion of ritual violence to include animals and objects; our often exclusive focus on human victims appears to have only perceived the tip of the proverbial iceberg. From these many new perceptions, more questions and agendas for future research crystalized. During the course of development, a few of the original speakers were unable to submit a manuscript as we moved forward with this

x vi   Preface x vi   Verso and Runninghead Acknowledgments

edited volume, and we wish to express our sincere thanks to Alana Cordy-Collins, Rose Tyson, Andrew Nelson, and Carol Mackey, whose papers directly influenced the thinking in this book in multiple ways. Each chapter was first commented on by up to three anonymous peer reviewers, and then by two outside reviewers selected by the University of Texas Press. We owe them all extensive debts of gratitude in helping to improve the quality of the chapters for publication. A 2012–2013 Utah Valley University Presidential Fellowship for Scholarly Projects awarded to Haagen Klaus assisted him in completing the first phase of editing and assembly of the volume, and a Spring 2015 faculty research leave at George Mason University assisted him in completing the book. At Utah Valley University, we thank Brian Birch, Kat Brown, Paul Bybee, Steven Clark, Daniel Fairbanks, Sam Rushforth, and David Yells. At George Mason University, we thank Amy Best, Andy Bickford, Deborah Boehm-Davis, Les Kurtz, and Daniel Temple. At the University of Central Florida, we acknowledge the support of Arlen Chase, Tosha Dupras, John Schultz, Joanna Mishtal, Beatriz ReyesFoster, and John Walker. Jennifer Dewey of the University of Central Florida ably compiled the bibliography. Haagen Klaus extends special recognition to Jenny Peterson for her valuable editorial assistance and patience during this project. Finally, we extend our gratitude to the editorial board at the University of Texas Press and the UT Press editorial staff, especially Casey Kittrell, Angelica Lopez, Victoria Davis, and Theresa May. Jay Marchand ably compiled the index for this volume Haagen D. Klaus, Fairfax, Virginia J. Marla Toyne, Orlando, Florida

Recto Runninghead  x vii

Ritual Vi o le n c e i n t h e A nc i en t An des

x vi i i   Verso Runninghead

Ritual Violence on the North Coast of Peru  1

chapter one

Ritual Violence on the North Coast of Peru Perspectives and Prospects in the Archaeology of Ancient Andean Sacrifice Haagen D. Klaus Department of Sociology and Anthropology, George Mason University

J. Marla Toyne Department of Anthropology, University of Central Florida

Throughout antiquity and into the present day, human behavior is remarkable for its complexity, spanning diverse peace-seeking strategies to a capacity to visit a disturbing range of harm and cruelty upon one another (Farmer 2003; Fry 2005, 2013; Knüsel and Smith 2014; Galtung 1969; Martin et al. 2012; Otto et al. 2006; Shulting and Fibiger 2012; Scheper-Hughes and Bourgois 2004). Within the continuum of violence, ritual violence is distinct. Encountered in almost every culture, it often transforms an otherwise profane act of murder into either a justifiable secular rite or an offering to the divine. Ritual violence may be associated with ideas surrounding veneration, supplication, destruction, transcendence, the value of human life, and personhood. Simultaneously, it may also embody deep reflections of politics, social organization, historical processes, and economics. Accordingly, the archaeology of ritual violence actively seeks to eschew simplistic or sensational notions and to open unique windows on ancient worlds that may be otherwise lost over time. Ritual violence was particularly varied and enduring in the South American Andes. This book aims to provide an initial regional synthesis on ritual killing for the pre-Hispanic north coast of Peru, where a unique sequence of independent human societies materialized over some ten

2   Klaus and Toyne

millennia. This book illustrates the development of distinct forms and diverse trajectories of ritual violence that developed during the final 1,800 years of prehistory while offering new concepts, methods, and interpretations of sacrifice on the north coast and beyond.

Theoretical and Conceptual Background: Definitional Complexities Early anthropological analyses that grappled with ritual violence, such as work by Tylor (1874) and Hubert and Mauss (1964[1898]), are perennially stimulating. A more recent definition proposed by Tatlock (2006) characterized sacrifice as an act of killing with the direct purpose to affect the realm of supernatural forces. These can be useful starting points to explore numerous expressions of ritual violence (Cucina and Tiesler 2007; López and Oliver 2010; Pluskowski 2012; Porter and Schwartz 2012; Tiesler and Cucina 2007; Tung 2012). Sacrifice is a complex form of human ritual, where intricate ideas are translated into a performance of symbolic communication. While archaeological evidence of sacrifice might be conceptualized in terms of a single event, sacrifice is a multifaceted social and ceremonial process—one preceded, paralleled, or followed by many linked actions. Further, many different perceptive vantage points can coexist among participants, witnesses, or victims of sacrifice; ritual killing may be understood differently even within a single society (Bell 1997). Sacrifice truly demands a broader perspective (Turner 1977; Valeri 1985). Accordingly, this introduction does not propose a single definition of sacrifice for the entire book. Instead, various chapters explore potential meanings, nuances, caveats, variations, and complexities of ritual killing within specific settings. Still, it is useful to consider that the word sacrifice, in its root Latin etymology, implies consecration or creation of sacred-ness through a connection to the divine. Girard (1977) holds that violence, in its capacity to end life, broadly represents the essence and “secret soul” of the sacred. A degree of violence appears universally involved when the offering (or oblation) is sacrificed (or immolated), regardless of generative or destructive motivations (i.e., Duncan [2005]). Cross-culturally, sacrifice has been seen as a highly specialized and ultimate means of communication affirming relationships between humans and supernatural forces (Girard 1977[1972]; Hubert and Mauss 1964[1898]; Oestigaard 2000). Life-ending violence might also be the price of exchange between

Ritual Violence on the North Coast of Peru  3

humans and the divine—a reciprocal act where something embodying the most precious value is given up in order to receive some benefit (Bloch 1992; Humphrey and Laidlaw 2007). There is often an implicit association between the sacrificer and the victim; the closer the killer is to the supplicant, the better (Hubert and Mauss 1964[1898]). The victim may be seen as a substitute for the individual conducting the sacrificial act, as autosacrifice would be the most powerful offering of all. Schwartz (2012) considers additional broad, cross-cultural themes linking sacrificial violence and consumption, where ritual provides powerful sustenance or nourishment to the divine (Trigger 2003). Also, cross-cultural practices frequently sacrificed animals, inanimate objects, plants, and liquids (McClymond 2008; Osborne 2004). Human sacrifice may only represent one part of a broader ceremonial spectrum. As a highly specialized form of symbolic communication, sacrificial activities tend to stand out from other forms of archaeological evidence of ritual (Marcus 2007). The nature of sacrifice promotes relatively repetitive, intentional, and internally consistent practices. When skeletal assemblages are consistently characterized by skewed age and sex variation found in a non-cemetery setting, ritual offering in some form might be involved. Deviation from modal funerary practices in terms of burial location, peri- and postmortem body treatment, unusual body or limb positioning, and use of associated grave goods (or lack thereof ) also potentially highlight some kind of specialized ritual framework (Eeckhout and Owens 2008; Shay 1985). Conversely, it is important to maintain vigilance regarding mortuary rituals that mimic sacrificial activities—not every unusual burial is necessarily a ritual killing. Archaeological evidence of sacrifice may be organized in terms of Schwartz’s (2012) classification of ritual killing, which identifies three broad conceptual categories. First, offering sacrifices involve supplication to the supernatural by ending the life of a living thing or the functional destruction of an inanimate object. Second, foundation (or construction) sacrifices are associated with acts of dedication, sanctification, or consecration with the offering medium placed into basal architecture, walls, or floors of ritual spaces or monumental structures. Third, retainer sacrifice is associated with high-status funerary contexts. This last category is challenging, as it can involve an offering without a sensu stricto sacrifice: ostensible “victims” may have either willingly committed ritual suicide to accompany high-status personages in death, or died before they were placed in a tomb.

4   Klaus and Toyne

Schwartz’s framework is indeed very useful, as many ancient ritual deaths may fall into one or more of these categories. It is simultaneously important to be on guard against etic and typological classifications that investigators might unintentionally invent. There may (or may not be) considerable conceptual overlap between these three broad categories of killing in an ancient society. It is also important to actively question when ritual killings may not be so easily labeled to challenge typological classification. Academic thinking, even today, sometimes reflects deeply rooted ideas about sacrifice as an impulse of a naïve or superstitious populace trying to satisfy a one-dimensional need (i.e., instigation of rainfall during a drought). Instead, it may be more productive to avoid monolithic views of human sacrifice as a form of obligatory payment to the divine and, instead, to identify the specific factors underlying individual acts of violence whenever possible. More nuanced frameworks might also attempt to consider how sacrifice tries to create a balance through which people attempt to commune with and negotiate with the supernatural and manipulate life, death, and their social world. It is also probable that acts of sacrifice involved multiple social agents that were at play with diverse priorities, values, perspectives, and practices. It is also possible that pre- and post-sacrifice activities held equal or greater significance than acts of killing. Ritual violence can also simultaneously embody vital secular and sociopolitical significance. Sacrifice may have held a special role in the emergence of hierarchical societies, where nascent elites made claims to a monopoly over life, death, and socioeconomic control through ritual (Morris 2007; Trigger 2003). Similarly, Fowler (1984) argues that human sacrifices in Mesoamerica functioned to consolidate power and strengthen the prestige and privilege of a ruling elite. In the Andes, the concept of lordship surrounding the office of the curaca involved lords who wielded ultimate power over life and death (Ramírez 1996). Many scholars have demonstrated that ritual violence in pre-Hispanic societies fulfilled religious functions while also serving more secular means of sociopolitical integration and control (Arnold and Hastorf 2008; Ramírez 2005). Other interpretations might see sacrifice as a tool of statecraft employed to terrorize a citizenry into submission. Ostentatious displays of ritual violence, however, were probably quite ineffective to achieve stable or lasting forms of sociopolitical integration (Yoffee

Ritual Violence on the North Coast of Peru  5

2005) and may only be associated with leadership tactics responding to significant short-term instability or insecurity (Carrasco 1999).

The Archaeology of Andean Sacrifice More than a dozen years ago, Elizabeth P. Benson and Anita G. Cook published Ritual Sacrifice in Ancient Peru (2001, University of Texas Press). Their edited volume spanned studies of art history, archaeological data, and human remains. Benson and Cook took initial steps toward a comprehensive treatise of sacrifice in the Andes. However, animal, vegetal, liquid, and object offerings were largely beyond the focus of their book, and chapters principally focused on Andean cultures possessing representational art styles. Chapters by Bourget and Verano provided particularly influential roadmaps guiding the following decade of sacrifice studies on the north coast. Verano especially showed how human remains provided a long-needed and independent perspective on Andean sacrifice. The North Coast of Peru: Natural Setting and Cultural Chronology A brief characterization of the natural setting and archaeological background of the north coast of Peru serves as the jumping-off point for the chapters in this book. The Central Andean region of northwest South America is one of the most ecologically and geographically diverse regions on Earth and was a cradle of singular cultural developments. The arid Central Andean coast corresponds to the 2,400-kilometer coastline of modern Peru. The combination of a warm equatorial landmass with cold coastal water temperatures has inhibited significant coastal rainfall since the terminal Late Pleistocene (Craig 1985). The desert coast is interspersed by some 40 roughly triangular valleys watered by rivers that drain from the Andean highlands into the Pacific Ocean. The north coast of Peru (figure 1.1) spans over 400 kilometers of coastline north of the modern capital of Lima. The 15 river valleys in this region, from north to south, are the Olmos, Motupe, La Leche, Lambayeque, Reque, Zaña, Jequetepeque, Chicama, Moche, Virú, Chao, Santa, Nepeña, Casma, and Huarmey valleys. Superficially, these coastal valleys could be seen as monotonous deserts, but each features a complex spectrum of microenvironments that helped sustain a long history of complex cultures.

6   Klaus and Toyne

1.1. A map of the north coast of Peru, highlighting the sites discussed throughout this book. Map by Haagen Klaus.

Ritual Violence on the North Coast of Peru  7

From First Settlement to the Growth of Complex Cultures: 9,000 BC to AD 100 The sequence of human societies in the Andes began some 11,000 years ago (figure 1.2). Early coastal inhabitants employed diverse approaches to hunting, fishing, gathering, maritime foraging, and limited gardening subsistence while occasionally venturing into the highlands in pursuit of new resources and exchanges of goods with neighbors (Dillehay 2011). The transition to sedentary life was slow, complex, and uneven. Around

1.2. The sequence of principal regional cultural developments on the coast of Peru. Drawing by Haagen Klaus.

8   Klaus and Toyne

3000 BC, a surge of supracommunal projects involved construction of monumental civic-ceremonial centers along the central and north coasts at sites such as Ventarrón, Huaca Lucía, Salinas de Chao, Aspero, El Paraíso, Caral, the Norte Chico region, and the gargantuan Sechín Alto (Alva Meneses 2012; Burger 1992a; Pozorski and Pozorski 2008; Shady 2009). After 1500 BC, the Cupisnique culture emerged on the north coast, with a core region in the Chicama, Jequetepeque, and Lambayeque valleys. Cupisnique society likely possessed a centralized, chiefdom-like structure that integrated diverse peoples on an inter-valley level (Elera 1998). Cupisnique religious art, while comparatively crude in light of later traditions, is captivating in its depictions of humans, stylized felines, and supernaturals. The Cupisnique decline around 600 BC is poorly understood, but they may have been destabilized by a major El Niño climate event and a possible tsunami (Elera 1998). The subsequent Salinar style emerged largely in the Moche, Virú, and Santa valleys. Though poorly studied, the Salinar possibly represented a coastal expansion of highland peoples that took advantage of the Cupisnique collapse. Strife may have developed between the intrusive Salinar and Cupisnique descendants, as some of the first fortifications on the north coast were constructed during this time (Willey 1953). As the Salinar declined, the Gallinazo (or Virú) culture rose to prominence around 200 BC. Diagnostic Gallinazo ceramic styles were present on nearly the entire north coast, but their strongest manifestations are noted in the lower Chicama, Moche, Virú, and Santa valleys. The Gallinazo chiefdoms constructed several notable adobe brick civicceremonial centers, including the iconic Gallinazo group in the Virú valley (Bennett 1939; Willey 1953). Recent advances in Gallinazo archaeology have clarified elements of their internal organization and legacy (Millaire and Morlion 2009). A key issue involves the fact that Gallinazo artistic conventions, construction techniques, mortuary patterns, and apparent genetic links to the later Moche (Donnan 2009; Shimada 2010) indicate distinctions among political, biological, and artistic features. Some Gallinazo polities likely evolved into Moche polities, whereas others existed alongside the Moche (Bourget 2003). The Gallinazo may be hypothesized as representing the roots of a north coast biological and cultural substratum that continued to exist, in varying forms, under the surface of later north coast societies (Donnan 2009; Klaus 2014a; Shimada and Maguiña 1994).

Ritual Violence on the North Coast of Peru  9

The Moche Culture: AD 100–750/800 The Moche represented the primary development on the north coast of Peru during the first millennium AD. It is perhaps now the most broadly examined pre-Inka Andean society (Benson 2012; Bourget and Jones 2008; Pillsbury 2001; Quilter and Castillo 2010; Uceda and Mujica 1994, 2003). Sometime around the first century AD, Moche Phase I–II (or the Early Moche period: AD 100–300) coalesced. Moche genesis probably occurred in multiple locations along the north coast owing to sociopolitical changes within the preexisting Gallinazo regional hierarchy. Early Moche elites were likely allied through gift-giving and shared religious beliefs and art styles (Shimada 2010). The spread or adoption of a Moche identity among non-elites likely represented a different process altogether (Bawden 2001:290–291) and, at the very least, united large numbers of fishers, farmers, camelid herders, and artisans into an integrated political economy. Moche origins also reveal the existence of a geopolitical and cultural bipartition within the north coast. The two principal loci of these early developments occurred between a prevailing northern polity centered around Sipán in the Lambayeque region and a secondary southern polity centered in the Moche-Chicama region (Castillo and Donnan 1994a; Shimada 1994b). By Moche Phase III–IV (or the Middle Moche era: AD 300–550), the regional balance of power shifted toward the southern MocheChicama polity. They eclipsed and incorporated the northern Moche while extending control southward as far as the Huarmey valley (Shimada 1999). The center of the southern Moche world was the site of Moche (or Huacas1 de Moche). There, the monumental Huaca del Sol (Temple of the Sun) was built from an estimated 140 million–plus handmade adobe bricks. Some 500 meters to the east was the principal religious center, the Huaca de la Luna (Temple of the Moon). An urban zone thrived between the two platforms and contained high-status residences, craft workshops, and cemeteries (Chapdelaine 2001; Uceda and Armas 1998). Variations in mortuary treatments and settlement patterns indicate Moche social organization involved a multi-tiered, hierarchical class system ruled by a small core of political and religious elites (Millaire 2002). At the zenith of Moche IV (AD 450–550), the southern Moche held sway over 350 kilometers of coastline. This may have represented the region’s first state (i.e., Billman 2002) or a confederation of closely

1 0  Klaus and Toyne

associated but competitive paramount chiefdoms (Shimada 1999). Others argue for a “many Moches” model (Quilter and Koons 2012), in which multiple decentralized non-state polities were realigned by a religious system into a new political economy and social configuration. An abnormally intense El Niño event contributed to various social and political transformations around AD 550 (Shimada et al. 1991). The southern Moche largely imploded. A new configuration of ideology emerged at Galindo in the middle Moche valley as elites distanced themselves from the recently discredited political-religious system (Bawden 2005). Farther north, a separate reorganization marked the beginning of Moche V (or the Late Moche period: AD 550–750/800). At the urban Lambayeque valley settlement of Pampa Grande, Shimada (1994a) argues, a state-level system emerged out of the Middle Moche collapse as the balance of geopolitical power swung back to the northern north coast. The final political disintegration of the northern Moche occurred around AD 750 (Shimada 1994a), though a few strongholds persisted until AD 800 or later (Castillo 2003; Uceda 2010). The Transitional Period then unfolded (ca. AD 750/800–900), and in some reaches along the southern north coast, this era lingered beyond AD 1100. This is perhaps one of the most regionally variable periods of north coast prehistory, and a highland Cajamarca presence may have thrived for a short time in the Lambayeque and Jequetepeque areas (Bernuy and Bernal 2008; Shimada 1994a). The Sicán (Lambayeque) Culture: AD 900–1375 From AD 800–900, the Sicán (or the Lambayeque Culture) ascended to regional power. By the Middle Sicán period (AD 900–1100), the religious-funerary complex of Sicán in the Pomac desert forest (La Leche valley) was the capital of a theocratic state holding influence from the Ecuadorian frontier to the Chicama valley (Shimada 1990, 2000, 2014a, 2014b, 2014c). The Middle Sicán period saw the development of a multiethnic society and a unique religious ideology revolving around an elite ancestor cult (Shimada 2014b). They advanced technological knowledge and metallurgy to bring Peru into the so-called Bronze Age while reviving monumental construction techniques (Shimada 2014a). Much of the non-elite citizenry appears drawn from the same local population that existed during Moche and Gallinazo times, and many inferred Moche descendants expressed an ethnic Muchik identity (Klaus 2014a)

Ritual Violence on the North Coast of Peru  11

expressed as a cultural substratum just under the surface of the dominant society. The Middle Sicán heartland of the La Leche–Lambayeque valleys supported large-scale irrigation agriculture, camelid pastoralism, marine resource exploitation, and bronze production (Shimada 1982) that underwrote a vast trade network. Middle Sicán administrators did not pursue militaristic subjugation of neighboring regions; rather, they apparently pursued alliances with local lords, corporate groups, and lineage heads within and beyond the Lambayeque heartland (Prieto 2010). Between AD 1050–1100, another powerful El Niño struck, followed by a 30-year drought. This was associated with the systematic torching of the temples atop the huacas at Sicán in a concerted effort to remove the leadership (Shimada 2000). During the Late Sicán (or Late Lambayeque) period (AD 1100–1375) there is little to suggest daily life changed outside of religion and politics. A new capital was established at Túcume, and Late Sicán influence continued into the northern Piura region but extended no farther south than the Jequetepeque valley. The Rise of the Chimú Empire and Conquest of the Inka: AD 900–1532 By the twelfth century AD, the regional balance of power was tipping back once again to favor the southern north coast as the Chimú kingdom began to expand out of the Moche valley. The El Niño–related disruptions that destroyed intra-valley irrigation networks ca. AD 1100 may have been an “initial spark” of Chimú expansion. Kolata (1990:135) envisions this shifted Chimú economy from provincial exploitation of local resources to a “parasitic extraction of foreign resources.” After the Chimú consolidated their core territory, a wave of expansion ca. AD 1200 reached the Jequetepeque region (Mackey 2006, 2011). The third phase of conquest around AD 1375 set the final boundaries of the empire from Ecuador in the north to the outskirts of modern Lima in the south (Mackey and Klymyshyn 1990). The Chimú capital of Chan Chan is the largest pre-Hispanic Andean city ever documented, covering nearly 25 square kilometers. Immense quantities of material wealth and human capital went into its construction, which consisted of nine hierarchically organized ciudadelas (monumental royal compounds, most likely built in pairs [Cavallaro 1997]), 35 smaller elite compounds, and a multitude of dispersed neighborhoods

1 2   Klaus and Toyne

(Day 1982). Toward the southern end of each ciudadela were massive burial platforms believed to function as a mausoleum for the dead king’s ancestor cult (Conrad 1982). The internal organization of the ciudadelas served practical administrative functions but also communicated unquestionable power and immutable order (Shimada 2000:97). The Chimú state appears largely secular, though Chimú artisans depicted a celestial deity that may have resynthesized earlier Moche, Middle Sicán, and Wari concepts. Water and ocean themes are very common in Chimú iconography and murals (Jackson 2004; Pillsbury 1996). Ceramics, typically monochrome dark gray wares, were mass-produced from simple two-piece molds and lacked careful craftsmanship. Inevitably, the Chimú came into contact with the Inka and refused Inka diplomatic overtures to be integrated into the empire of Tiwantinsuyu. According to Cabello Balboa (1586 [1751]), Inka armies brutally routed the Chimú around AD 1460 or 1470. One of the most significant impacts of Inka conquest involved land tenure aimed at centrally controlled production (Ramírez 1990:519–525). The Inka sought to control the coastal trade system and began to convert ceremonial events into opportunities for barter and trade that would become tribute (Ramírez 1990:532). As the Chimú before them, the Inka were focused on managing the uppermost levels of political and economic structures, not daily comportment or beliefs. Furthermore, the Inka also encountered a highly organized north coast Chimú society whose deeply entrenched customs could not be changed overnight (Ramírez 1990:532). A Review of North Coast Sacrifice Sacrifice is a relatively new addition to the topics and motivations underscoring scientific Andean archaeology since the late nineteenth century (Shimada and Vega-Centeno 2011; Tanteleán 2014). Numerous ethnohistoric sources from the sixteenth to eighteenth centuries recount mostly Inka ritual killing in the central and southern Andean highlands (Besom 2009; Duviols 1976). Archaeological evidence of sacrifice first emerged from Uhle’s (1991[1903]) late nineteenth century excavations at the coastal oracle center of Pachacamac, where a group of Inka-era females were found with neck ligatures. Yet 83 years passed until a modern focus on sacrifice began to emerge through the development of contemporary physical anthropology and groundbreaking work of John Verano (1986a).

Ritual Violence on the North Coast of Peru  13

Ethnohistory and the Inka Legacy in Scholarly Research Ethnohistory has long been a conceptual starting point in the study of Andean sacrifice and identifies different forms and intents of Inka ritual violence. Qhapac hucha (capacocha or capac hucha) sacrifices often involved mountaintop offerings of children. The term is varyingly translated from the Quechua as “solemn sacrifice,” “royal sin,” “royal transgression,” “royal obligation,” or, in Aymara, “great sacrifice” (Gentile 1996). Qhapac hucha has been interpreted as a substitution sacrifice meant to embody the sins of the Inka (Classen 1993; Zuidema 1977– 1978). Runa killings entailed sacrifices of low-status, able-bodied males and warriors, while necropampa rites produced tomb retainers (Besom 2009). These sources are informative, provocative, and of great value. They are also problematic—drawn from memory or oral histories—and are strongly skewed toward Inka ritual. Regarding works by European chroniclers, readers must consider misperception, conflation, and likely misunderstanding of indigenous ideas (Ramírez 1996; Reinhard and Ceruti 2010). The chronicles were also written under various circumstances at different times and pursued multiple agendas. For instance, some European chroniclers justified their imagined racial and moral superiority by highlighting “barbaric” Inka beliefs (Cobo 1990 [1653]). Garcilaso de la Vega (1945 [1609]), a descendant of the Inka royal family, conversely downplayed sacrifice to portray Andeans as a noble people to an ethnocentric European audience. Fieldwork by Andrushko et al. (2011), Ceruti (2003), and Reinhard and Ceruti (2010) has documented Inka sacrificial practices and corroborated some ethnohistoric observations. Remarkable information has been gained from these mountaintop contexts containing astonishingly well-preserved victims in the highest archaeological sites on Earth. Victims often wore high-quality textiles and garments, including headdresses reflecting honorary status, and were provisioned with Spondylus shells, food, and coca leaves (Ceruti 2004). Stable isotope variation in their hair demonstrates a possible ritual sequence that followed victim selection where antemortem preparation included long-term fêting and a shift to a distinctive ritual diet before movement across geographic space preceding death (Fernández et al. 1999; Wilson et al. 2007). Some supplicants were dispatched via strangulation, drowning, or a blow to the head. Others appear to have been entombed alive and to have perished

14   Klaus and Toyne

from starvation, freezing temperatures, the effects of high altitude, or a combination of all three (Reinhard and Ceruti 2010:125–126). When contextualized, the ethnohistory and archaeology of Inka practices cover only a brief span of just one Andean culture. They may help us to understand the archaeology of other traditions but cannot be taken as representative of an essentialist “Andean” kind of sacrifice via lo andino (or pan-Andean) analogies. Roe (2008:189) considers this fallacy as a kind of modern Inka “tyranny,” whereby investigators look at remote prehistory through the lens of the last major society before European contact and mistakenly collapse a vast diachronic diversity into a conflated synchronic reality. Roe has a point, and here we seek to first characterize north coast rituals and cultures on their own terms. Only then may reliable comparison with other societies be feasible. Sacrifice Among Early North Coast Cultures: 7800 BC to AD 100 A specialized engagement with the dead likely existed among the early inhabitants of the Zaña valley ca. 7800–5800 BC. There, studies of human remains have identified potential signs of ritual cannibalism (Rossen 2011; Verano and Rossen 2011). Links to ritual violence seem improbable, as these assemblages are reasonably interpreted as reflections of mortuary rituals and garden magic associated with the transition from foraging to farming. Understanding of ritual killing in later Preceramic and Formative cultures is also limited. However, on the central coast of Peru, an elaborate burial of a child in a temple at Aspero (post–3000 BC; Feldman 1980) possesses the hallmarks of a foundation sacrifice, as do other child burials at Paloma (5000–2800 BC; Quilter 1989) and Ancón (ca. 400 BC; Burger 1992a). In the north highlands, two children were found under structures at Alto Huallaga (Kotosh Chavín phase, 700–250 BC; Onuki 1993). While sacrifice is certainly possible, there is no evidence of trauma. It cannot be determined with certainty that such children were violently dispatched. Other explanations, including natural death and symbolisms involving offering without sacrifice, must be considered. Such contexts may forever be ambiguous. Furthermore, the immolation of objects, rather than people, seems widespread during the Initial and Early Horizon Periods. The highland Mito architectural tradition (ca. 3000–1800 BC) involved construction of hearths within temples in which material offerings of some kind were

Ritual Violence on the North Coast of Peru  15

burned (Bonnier 1997). Comparable fire-mediated offering/transformation/destruction has been documented at Ventarrón in the Lambayeque region (ca. 2500 BC; Alva Meneses 2012) and with the “temple entombment” of Huaca Lucía in the Pomac forest between 600–550 BC (Shimada 1986). Elsewhere, at Chavín de Huantar in the adjacent north highlands, rituals involved intentional destruction of ceramic vessels within subterranean canals and various galleries (Kaulicke 1991). The manner in which these objects were systematically deposited or smashed in situ is powerfully consistent with object immolation (Rick 2013). By the time of the Cupisnique (1500–650 BC), religious iconography entwined images of human decapitation with metaphors of vegetal fertility, cuttings of root plants, and liquid nourishment (Elera 1998; Jones 2010). Cupisnique artisans depicted five Decapitators that harvested human heads (Cordy-Collins 1992) (figure 1.3). One sculpted ceramic bottle graphically depicts an individual having slit his own throat in an ostensible depiction of auto-sacrifice (Cordy-Collins 2001a). While other contemporaneous coastal peer polities depicted disembodied and bleeding human heads in murals, none surpass the carved stone reliefs at Cerro Sechín (lower Casma valley). There, warriors are portrayed between depictions of mutilated human bodies and body parts such as bleeding, decapitated heads and piles of eyeballs and vertebrae (Samaniego et al. 1995). Yet, it is unknown if Cupisnique peoples or their contemporaries conducted ritual killing in actuality. Physical evidence of sacrifice continues to be absent, and we must caution against potential disjunctions between art and practice. However, one candidate sacrifice site is the Late Cupisnique seaside temple of Morro de Eten (Reque river valley) (Elera 1986). There, erosion and looting exposed skeletal remains from under temple floors atop the small mountain. These several individuals bear a resemblance to definitive sacrifices at the adjacent Cerro Cerrillos (Klaus et al. 2010). The windswept summits of Morro de Eten require a new round of large-scale excavation, but, realistically, damage to the site may have destroyed whatever evidence might have once existed there. Ritual Violence in Moche Visual Art and Practice: AD 100–750/800 For decades, Moche iconography was the center of gravity in the study of north coast sacrifice. An extraordinary repertoire of fineline paintings

1 6   Klaus and Toyne

A

b

c

1.3. The Cupisnique iconographic corpus contained five distinct Decapitator entities, including: (A) a stylized human, (B) arachnids, and (C) fish. These particular images were all carved on Cupisnique stone bowls or vases. Redrawn by Haagen Klaus from Cordy-Collins (1992: figures 2, 5, and 6).

adorning ceramic funerary bottles and large-scale painted murals depicted Decapitators (figure 1.4), dueling warriors (figure 1.5), and the taking of captives (figure 1.6) (Cordy-Collins 1992; Donnan 1978; Donnan and McClelland 1999). Such motifs have been systematically linked to a broader sequential narrative of prisoner-taking, the ritual killing of deer and sea lions, and human sacrifice (Bourget 2001a; Cordy-Collins

Ritual Violence on the North Coast of Peru  17

1993; Donnan 1997a; Verano 2001b). At the heart of the narrative, Alva and Donnan (1993: fig. 143) identified the “Presentation Theme” or “Sacrifice Ceremony” (figure 1.7). Produced during the Moche IV period by southern Moche artisans, this scene was the culmination of various activities where human blood sacrifice was consummated. In 1995, Steve Bourget’s excavations at the Huaca de la Luna’s Plaza 3A uncovered the remains of victims of human sacrifice (Bourget 1998a, 1998b, 2001a, 2001b; Bourget and Newman 1998). Analysis demonstrated that injuries, body treatments, and victim identities largely paralleled depictions in Moche iconography (Verano 1998, 2001b, 2008a;

A

b

c 1.4. Moche Decapitators. (A) An early Moche Phase I/II fineline painting of a spider Decapitator adorning a stirrup-spout ceramic bottle. Redrawn by Haagen Klaus from Donnan and McClelland (1999: figure 2.20); (B) An early Middle Moche spider Decapitator, of the style adorning the metal backflaps and gilded copper bells of the Lord of Sipán (redrawn by Haagen Klaus from Alva and Donnan 1993: figure 152); (C) a late Middle Moche (Moche IV) large-scale polychrome adobe mural at Huaca de la Luna’s Great Patio, Platform I, Structure D, Lower Level. Photo by Haagen Klaus.

1 8   Klaus and Toyne

1.5. Rollout drawings of combat scenes on Middle Moche ceramics that depict Moche warriors engaged in hand-to-hand combat. Redrawn by Haagen Klaus from Donnan and McClelland (1999: figures 4.1 and 4.24).

1.6. Rollout drawing of fineline paintings show captives depicted as stripped nude, sometimes bleeding from their faces, and with a rope around their necks. Redrawn by Haagen Klaus from Donnan and McClelland (1999: figures 6.22 and 6.23).

d

A b

c

e

f

1.7. Rollout drawing of the Moche IV Presentation Theme, or the Sacrifice Ceremony. On the bottom register, an anthropomorphized feline (E) and human priestess (F) slit the throats of nude and bound captives. On the top register, the principal figure (A) receives a goblet, inferred to contain sacrificial blood, from two priests (B, D), a priestess (C), and attendants. Redrawn by Haagen Klaus from Donnan and McClelland (1999: figure 4.102).

Ritual Violence on the North Coast of Peru  19

Sutter and Verano 2007; Toyne et al. 2014). This demonstrated that Moche art realistically depicted actual rituals. Subsequently, additional examples of Moche sacrifice have been identified at Huaca de la Luna’s earlier Plaza 3C (Tufinio 2008; Verano 2008a). In light of these finds, Uhle’s (2014 [1913]) excavations atop Cerro Blanco (the mountain abutting the east side of Huaca de la Luna) that documented incomplete human skeletons and isolated crania intensely evoke the Moche mountain sacrifice theme (Benson 2012: fig. 10.8). Choreographed foundation burials of children were also found at Huaca de la Luna Plaza 3A but, again, lack evidence of violent death (Bourget 2001a). To the north, a possible Moche sacrifice victim was found atop Huaca Cao Viejo (El Brujo Complex, Chicama valley) (John Verano 2014: personal communication). On the north face of the Cao Viejo mound, the proximal femur of a mutilated individual was intentionally placed within a large-scale mural depicting prisoner-taking (Verano and Lombardi 1999). A cache of decapitated heads and evidence of decapitated dedicatory burials were identified at Huaca Dos Cabezas (Jequetepeque valley) (Cordy-Collins 2001a; Klaus 2013). In the south, Gagné (2009) studied 18 sacrificed individuals at El Castillo de Santa, corresponding to a Gallinazo peer polity in the Santa valley during the Moche era. There, practices of throat-slitting, blunt force cranial trauma, decapitation, and complex patterns of dismemberment were intriguingly similar to that observed at Huaca de la Luna, further blurring the lines between Gallinazo and Moche cultural phenomena. During the Late Moche era, sacrifice is far less frequently depicted outside of a small thematic component of the Late Moche “Burial Theme” involving a female victim (Donnan and McClelland 1979; Bourget 2006). Huaca de la Luna’s Late Moche Platform III contains evidence of chest mutilation or chest opening at the New Temple of Platform III (Verano and Backo, in press) and may even have earlier roots at Plaza 3C (Hamilton, chapter 2). Swenson and Warner (2012) identified probable foundation sacrifices of two women at the Late Moche site of Huaca Colorada (Jequetepeque valley). A double burial in a Late Moche temple at the Ventarrón complex (Reque valley) contained the headless body of an adult (seemingly decapitated) accompanied by several ceramic bottles (Alva Meneses 2012). One of the vessels was a representation of a Decapitator, very carefully broken so as to remove the ceramic Decapitator’s head.

2 0  Klaus and Toyne

Late Pre-Hispanic Sacrifice: AD 900–1532 By the Late Intermediate Period, representational art styles waned yet further, and it was sometimes assumed that Moche-style sacrifice similarly faded. However, within the core of Middle and Late Sicán society, killings occurred at Cerro Cerrillos, Cascajales, and Túcume; while these rituals had precedents partially rooted in earlier Moche practices, they also involved new ideas (Klaus 2014b; Klaus et al. 2010; Toyne 2011a, 2015a, 2015b). Additionally, Sicán ritual strangulation was identified at El Brujo (Verano 2001a). Although the Chimú were not known for representational art, at least one known stirrup-spout vessel (probably looted from the Moche/ Chicama region) depicts sacrifice. A human sacrificer pins a victim down while pulling his head back by the hair with his left hand. He is then poised to slit the victim’s throat with a knife held in the right hand (figure 1.8). Known Chimú practices bore similarities to this depiction. A mass grave dating to the provincial Sicán or Chimú occupation at Pacatnamú (Jequetepeque valley) contained the remains of young and middle-aged men (Verano 1986a) whose throats were slit and chests opened. A similar picture emerged from the discovery of inferred reprisal killings (not sacrifice) involving the mass execution of nearly 180 boys and men on a beach at Punta Lobos (Huarmey valley) (Verano 2007; Verano and Toyne 2011). Radiocarbon dates associate these deaths with the southward expansion of the Chimú Empire. More religiously minded activities were described by Donnan and Foote (1979) in inferred Chimú sacrifices overlooking the Pacific Ocean at Huanchaco (Moche valley). There, 17 subadults with possible perimortem cranial trauma were interred with camelids in shallow pits. Given the similarity of these contexts to a large-scale mass sacrifice of pairs of children and juvenile camelids found at nearby Huanchaquito (Prieto et al. 2014), it seems likely that these beaches represented a ritual locus just outside the Chimú capital of Chan Chan. A focus on sacrifice also persisted at Túcume, where performances of throat-slitting, chest-opening, and decapitation of children and adult men (along with camelid and object offerings) persisted under foreign Chimú and Inka regimes (Toyne 2011b).

Ritual Violence on the North Coast of Peru  2 1

1.8. A rare Chimú sculptural bottle depicting an act of throat-slitting is clearly reminiscent of earlier Moche motifs. Here, a human sacrificer has positioned himself atop a victim. He pins the victim down while pulling his head back by the hair with his left hand and is poised to slit his throat with a knife held in the right hand. Courtesy of the Museo Arqueológico Rafael Larco Herrera; photo by Haagen Klaus.

Current Questions While we may be confident that past work provides a basic characterization of sacrifice on the north coast of Peru after 30 years of research, we are just at the beginning of a process to assess the broader patterns of ancient ritual killing in the region. More questions than answers have emerged. Within what kinds of social and ecological contexts did people sacrifice one another? What kinds of ritual and symbolic unity or variations existed within and among societies and regions? In what manners did the act of killing articulate within broader pre-, peri-, and postimmolation rituals, such as feasting, acts of commemoration, veneration, or desecration? With the emergence of contextual, biocultural, and

2 2   Klaus and Toyne

social orientations in bioarchaeology (see chapters in Agarwal and Glencross [2011]), it is also possible to further examine sacrifice in terms of ancient personhood, social constructions of individual and group identities, the phenomenology and sensory experiences of violence, and relationships with mortuary rituals. Although the focus on human remains has been long overdue, how can we also enhance the emergent conceptualization and study of animal, object, and liquid offerings? The relative absence of evidence associated with organized warfare or endemic violence on the north coast of Peru (Klaus 2014b) stands in contrast to other Andean regions and times (Tung 2012; Arkush and Tung 2013; Kurin 2014). Was culturally sanctioned violence channeled into a ritual outlet or manipulated into an equalizer of social conflicts? If lords claimed power over life and death through sacrifice, might have some of their subalterns done the same, particularly in societies that contained diverse and competitive moieties? When might sacrifice represent a ritual “third space” (Bhabba 1994) where political dispute, social opposition, and other tensions were embodied, enacted, and resolved? This collection of research, through multiple paths and approaches, seeks to begin addressing such questions through four linked themes: (1) historical development and regional variation of north coast sacrificial rituals from the early first millennium AD to the European conquest; (2) initial definition of a continuum of ritual violence spanning people, animals, and objects; (3) establishment of evidence demonstrating pre-, peri-, and post-sacrificial ritual activities; and (4) integration of empirical methods with interpretive, contextual approaches to reconstruct ancient identities, behaviors, and beliefs. In this book, contributors demonstrate a roughly 1,800-year sequence of ritual killing that was, in many ways, quite distinct from that of the Inka. Others confront so-called gray zones where sacrifice is strongly suspected but evidence is ambiguous. Yet other contributors wrestle with defining physical and contextual signatures of religious violence versus secular punishments or executions. Various contributions weigh lo andino (pan-Andean) notions. The 15 chapters are organized into four sections: Part I: Ancient Ritual Variation and Methodological Advances in Studies of Sacrifice; Part II: Ancient Identities, Ambiguous Deaths, and Complex Burials; Part III: Continuums of Killing: Sacrifice of Animals and Objects; and Part IV: Perspectives from Beyond the North Coast of Peru. Each part opens with a short introduction discussing the chapters contained therein.

Ritual Violence on the North Coast of Peru  23

Conclusion and Future Prospects The reconstruction of ritual violence on the north coast of Peru has only recently commenced. Future discoveries—some of which we are incapable of anticipating today—will no doubt prompt us to reflect upon, revise, or reject many current ideas and interpretations. Still, the chapters in this book take important steps toward establishing coherent regional and temporal understandings of north coast ritual killing to highlight previously unseen variations of sacrifice on various spatial and temporal scales. The studies also underscore the ways in which the living interacted with sacrificed bodies, as victim burial shows both connections and discontinuities with known funerary behaviors. The emerging focus on ritual killing of animals (e.g., Goepfert 2010, 2012) and objects (Millaire, chapter 13) instructs us not to privilege human sacrifice and helps us expand current thinking regarding the ancient ideas surrounding the goals and intents of sacrifice. The studies in this volume also seek to illustrate that archaeological investigations into sacrifice require a diverse and creative range of analytical approaches. The authors respond to a point made earlier in this chapter: the study of ritual killing demands a broader perspective. Though much of this work is strongly rooted within empirical frameworks, the indispensable and complementary value of interpretive, iconographic, and theoretical orientations is clear. In other words, an ontological approach that seeks holistic understanding holds particular value for this kind of work. One key element is missing: the origins of ritual killing on the north coast of Peru. Currently, no one knows precisely where, when, and under what conditions traditions of sacrifice first began. This is partially a function of the concentrated study of first and second millennium AD north coast societies where traditions of sacrifice were already entrenched. Also, many sacrificial contexts have been encountered by blind luck in the study of an architectural complex or a rescue excavation. It might well be that the earliest sacrifices were probably not people but objects instead, such as those that we infer were destroyed in ritual offering fires at Ventarrón or Kotosh. Origins of animal sacrifice are even less clear. If the few dedicatory human burials of the Formative era were indeed sacrifices, the lack of trauma shows that the spilling of blood was not part of the early sacrificial liturgies. Truly, study of the

2 4   Klaus and Toyne

origins of ritual sacrifice on the north coast of Peru is long overdue and requires explicitly designed investigation. We suspect that resolution to these questions will be found buried somewhere within Cupisnique, Salinar, and Gallinazo sites. In closing, we hope this up-to-date and contextual synthesis of the current generation of archaeological studies of sacrifice on the north coast of Peru will foster nuanced, scientific, and humanized approaches to ritual violence. In addressing current unknowns and proposing many new questions, this book can contribute a better understanding of the ways in which Andean peoples interacted with their world through acts of sanctified violence. Notes 1. The term huaca (sometimes alternatively spelled wak’a or guaca) is derived from the highland Quechua language and denotes an object, place, mountain, animal, shrine, or person (truly, anything) that is imbued with sacredness or supernatural potency. Since the early days of Colonial Peru, the remains of pre-Hispanic monumental constructions were also referred to as huacas. In modern archaeological practice, huaca can be part of an archaeological place-name (i.e., Huacas de Moche) or part of the name of a specific pre-Hispanic structure within a larger site (i.e., the various named truncated platform mounds at sites such as Huacas de Moche, El Brujo, Sicán, and Túcume, to name just a few).

Ritual Killing, Mutilation, and Dismemberment  2 9

chapter two

Ritual Killing, Mutilation, and Dismemberment at Huaca de la Luna Sharp Force Trauma Among Moche Sacrifice Victims in Plazas 3A and 3C Laurel S. Hamilton Department of Anthropology, Tulane University

The Moche were a pre-Hispanic Andean culture that flourished on the north coast of Peru from AD 100–800. Their achievements included large-scale monumental architecture, resplendent metallurgy, complex political organization, and an art style unequaled in the prehistoric Western Hemisphere (Bawden 1996; Benson 2012; Bourget and Jones 2008; Shimada 1994a; Uceda and Mujica 1994). Art historians and archaeologists long interpreted artistic depictions of prisoner capture, mutilation, sacrifice, and dismemberment as part of a Moche mythic narrative (Donnan 1978). Yet, since the late 1980s, archaeological discoveries indicated that Moche iconography realistically depicted many of their rituals (Alva and Donnan 1993; Bourget 2001a, 2001b; Verano 2000, 2001a, 2001b). This chapter seeks to open new perspectives on Moche ritual killing via integration of macroscopic and microscopic analyses of cut mark patterning among human sacrifice victims excavated at the site of Huaca de la Luna. This study compares empirical evidence to Moche representations of sacrifice to specifically explore relationships among the location of cut marks on victims, the sacrifice process, and how it changed over time.

3 0  Hamilton

The Scenes and Settings of Moche Ritual Killing On the Peruvian north coast, macabre iconographic representations of ritual death and dismemberment appeared frequently in Moche art—also well known for its realistic portraiture, naturalistic styles, and representative nature. Tombs at Sipán (Alva and Donnan 1993), Úcupe (Bourget et al. 2012), and San José de Moro (Castillo and Donnan 1994a, 1994b) contained the remains of high-status individuals evidently matching various principal figures in the visual narrative of the Sacrifice Ceremony—reflecting the extent to which Moche artisans sought to represent their world. Other components of Moche art blended realism with the supernatural as they institutionalized elements of violence, death, and sacrifice. The Decapitation Theme (Cordy-Collins 1992, 2001a) typically featured a fearsome supernatural Decapitator entity holding a crescentbladed knife or tumi in one hand and a human head in the other (Alva and Donnan 1993; Franco et al. 1994; Lapiner 1976; Larco Hoyle 2001a, 2001b; Uceda 2001). Decapitators were alternatively portrayed as holding a tumi in one hand and grasping the soon-to-be victim by the hair with the other (Donnan and McClelland 1999), or with the tumi at or near the throat of the victim (Cordy-Collins 2001a; Donnan 1976). Decapitation and dismemberment were often associated with Moche contexts of combat between warriors. Moser (1974) and Bourget (2006) consider the primary objective of Moche warfare was not to kill an enemy but to capture an opponent for sacrifice. Donnan and McClelland’s (1999) “The Warrior Narrative” involved a sequence of scenes involving combat, capture, bleeding, parading, presentation, and sacrifice of prisoners (Alva and Donnan 1993; Castillo 2000a; Hocquenghem 1987; Verano 2001a, 2001b). On the north façade of Huaca Cao Viejo at El Brujo, a brightly painted mural depicts a Moche warrior leading a procession of nude prisoners linked together at the neck by a long rope while other warriors follow behind (Gálvez and Briceño 2001: figs. 15, 17, 18). After it was completed, deep incisions were made on the penises and legs of some prisoner images suggesting metaphors of torture, mutilation, or bloodletting (Gálvez and Briceño 2001). Other Moche ceramics depict nude male captives surrounded by severed heads, arms, and legs (Donnan and McClelland 1999). The Moche also crafted three-dimensional

Ritual Killing, Mutilation, and Dismemberment  31

ceramic representations of body parts, such as disembodied forearms with hands and lower legs with feet (Lapiner 1976; Purin 1990) perhaps symbolizing trophies (Benson 2001). The central activity linked to the Warrior Narrative is the Sacrifice Ceremony (or Presentation Theme) (Donnan 1978:158–173). This was a common motif among the southern Moche (especially in the ChicamaMoche river valleys) during the Moche IV era. The Sacrifice Ceremony appeared in fineline paintings, low-relief ceramics, and on a large-scale polychrome mural at Pañamarca (Nepeña valley) (Donnan 1976). Here, three warrior-priests and a priestess oversee a ritual in which bound and nude prisoners have their throats cut by blade-wielding supernatural or anthropomorphized attendants with the blood captured in ceremonial goblets (Donnan 1976). Throughout the 1990s, various archaeological discoveries on the north coast of Peru have demonstrated the reality of Moche human sacrifice, which in many cases appears to generally parallel their art with examples of throat-slitting, dismemberment, and decapitation (Alva and Donnan 1993; Bourget 2001b; Cordy-Collins 2001a; Verano 2001a, 2007). Still, many complex questions persist regarding the nature, structure, and variability of Moche ritual killing. Accordingly, in this chapter, four linked issues are examined through cut mark patterning at Huaca de la Luna. First, do frequency and location of cut marks show that victims were treated in a systematic and consistent manner within or between sacrificial contexts? That is, can an integrated macroscopic and microscopic analysis of cut marks identify and distinguish signature patterns of different activities? Second, what can macroscopic and microscopic studies of cut mark form and cross-sectional morphology reveal about the types of tools used by Moche ritual specialists to kill their victims? Are these wounds consistent with the use of a thin-bladed copper knife (i.e., a tumi), as has been long inferred, or some other type of metal or lithic tool? Also, what can microscopic analysis of cut mark directionality and handedness reveal about the positioning of the sacrificer relative to the victim? That is, how did the ritual specialists physically conduct these killings? Third, what does a diachronic comparison of cut mark patterns reveal about the evolution of Moche sacrifice at the Huaca de la Luna over time? Fourth, how concordant are cut mark patterns with sacrificial imagery as represented by Moche artisans?

3 2   Hamilton

Materials and Methods Modeling Mutilation: Behavioral Expectations from Sharp Force Trauma Studies Research in ethnoarchaeology and forensic anthropology has documented the macro- and microscopic characteristics of cut marks involving morphology, frequency, location, and orientation. These data can be applied to reconstruct and model perimortem and postmortem activities. Behavior and Intent via Cut Mark Patterns Decapitation signatures involve complete removal of the head via sharp force trauma near the base of the cranium and on specific aspects of the cervical vertebrae. Related cut marks on the skull are often located on or around the occipital condyles and the supra-occipital and basio-occipital regions of the cranium (Melbye and Fairgrieve 1994; Olsen and Shipman 1994). Decapitation may also produce multiple cut marks on the first cervical (C1) vertebra and on the dens and body of the second cervical (C2) vertebra (Melbye and Fairgrieve 1994; Verano 1986a), in addition to laminae, pedicles, spinous processes, and articular facets of intervertebral joints (Melbye and Fairgrieve 1994; Pijoan and Mansilla 1997; Pijoan and Lizarraga 2004; Verano 1986a, 2001a, 2008a). Although the first and second cervical vertebrae are most commonly affected in decapitations, cut marks may be observed on lower cervical vertebrae (Bush and Stirland 1991; Frayer 1997). Bilateral cut marks on the superior aspect of the first rib, anterior and superior aspects of the clavicle, and the first thoracic vertebrae indicate the head was severed at the base of the neck (Harman et al. 1981; Smith 1993) but can also be associated with throat-slitting. Occasionally, the mandible may be cut, especially along the posterior border of the ascending ramus, the gonial angle, and the inferior border of the mandibular corpus (Melbye and Fairgrieve 1994). Olsen and Shipman (1994:380) define the correlates of defleshing as “short fine cut marks or broader scraping over the surfaces of bones caused when a sharp tool is used to remove soft tissue adhering to the bone.” If the objective is the removal of large pieces of skin or muscle from a body, defleshing marks are primarily found in areas of muscle attachment (i.e., the spine and body of the scapula, the ribs, the laminae and spinous processes of the thoracic and lumbar vertebrae, the ilium,

Ritual Killing, Mutilation, and Dismemberment  33

and the diaphyses of long bones) (Fernández-Jalvo et al. 1999; Jelínek 1993; Lambert et al. 2000; Turner and Turner 1999; White 1992). If the goal is to separate the bones from soft tissue, defleshing marks may be less specifically patterned, occurring wherever skin, muscles, tendons, ligaments, periosteum, or any other soft tissue attaches to bone (i.e., the face and scalp region, the body and ascending ramus of the mandible, the clavicle, and the hands and feet) (Drusini and Baraybar 1991; Jamieson 1983; Massey and Steele 1997; Pijoan et al. 1989; Raemsch 1993; Verano et al. 1999). These two activities are not mutually exclusive and can occur in tandem. Defleshing marks typically appear in clusters, as repeated strokes often are necessary to completely detach muscles or remove soft tissue from bone. On limb bones, cut mark clusters frequently exhibit the same orientation (Olsen and Shipman 1994). In cases of intentional dismemberment, a sharp tool is used to reduce a body into smaller, constituent parts by cutting apart joint articulations (Olsen and Shipman 1994; Raemsch 1993; Turner and Turner 1999; Verano 2001b, 2008a). Locations frequently targeted by dismemberment activities include the area around the atlanto-occipital joint, mandibular condyles and coronoid processes, intervertebral joints, glenohumeral joint, condylar region of the humerus, radial neck, acetabulum of the pelvis, and femoral neck (Blom et al. 2003; Jelínek 1993; Pijoan and Pastrana 1987; Verano et al. 2007). Identification of Tool Types Experimental studies by Walker and Long (1977), Binford (1981), and Shipman and Rose (1983:64) demonstrated that macro- and microscopic techniques can distinguish between metal and stone tools via crosssectional morphology of the grooves left by the tool on the bone. Metal blade cuts appear as narrow, V-shaped grooves with smooth, straight sides that terminate in a distinct apex at the bottom of the groove (Walker and Long 1977:608). Binford (1981:105) noted that metal implements usually produce narrow (“almost hairline”), straight, and relatively long cut marks, and that metal tools often incise bone at an oblique angle, leaving behind a small overlapping bone shelf. In contrast, bifacially flaked stone tools produce wide and irregular cross-sections with concave sides that usually do not converge in a single apex. That is, they are coarser and possess U-shaped cross-sections. Stone tools typically leave groups of short, parallel marks on the bone

3 4   Hamilton

that are wider and “more ragged,” reflecting their thicker and irregular cutting edges (Binford 1981). Further, Shipman and Rose (1983:64) observed two diagnostic features of stone tool cut marks: shoulder effects and barbs. Shoulder effects are one or more short marks that originate from, and are made with, the same stroke as the main cut mark and either diverge from or run parallel to the main mark for some portion of its length. These result from “contact between the tool’s shoulder and the bone during cutting” and are produced by asymmetrical hand pressure. Barbs are probably caused by “small inadvertent motions of the hand either in initiating or terminating a stroke” (Shipman and Rose 1983:66). Directionality of Cutting Microscopic studies of cutting-stroke direction on bone were first undertaken as a means of obtaining data on butchery patterns, carcass processing techniques, and handedness of early hominins (Bromage 1987; Bromage and Boyde 1984; Shipman and Rose 1983). Bromage and Boyde (1984) presented three microscopic features—bone smears, oblique faulting, and oblique chipping—that they argued were reliable and consistent indicators of directionality. Bone smears, which under a scanning electron microscope (SEM) appear as small nicks in the floors, walls, and shoulders of the main groove of a cut mark, lift up in the opposite direction of the cutting stoke. Oblique faults are located on the surface of cut mark margins and intermittently cross the main groove. They separate short segments of bone that are displaced in the same direction as the cutting motion. Oblique chipping is described as “a characteristic pattern of breakage on one or both sides (often ‘V’ shaped) forward and away from the center of the mark in the direction of the cutting movement” (Bromage and Boyde 1984:363). However, these criteria do not occur in all cut marks. Multiple factors may affect their formation, including the cutting angle, shape of the cutting edge, variations in applied pressure, collagen fiber orientation, bone density, and idiosyncrasies of the tool user. Still, when at least one of the three criteria is present, Bromage and Boyde (1984) argue that directionality can still be accurately determined. The Archaeological Site of Huaca de la Luna Huaca de la Luna is one of two monumental structures that dominate the site of Huacas de Moche (or Pyramids of Moche), which was the center

Ritual Killing, Mutilation, and Dismemberment  35

2.1. The western façade of the Huaca de la Luna in 2014 with the mountain of Cerro Blanco overlooking the complex. Photo by Haagen Klaus.

of the southern Moche world at the peak of their civilization between AD 350–550. Huaca de la Luna appears to have predominantly served the religious elite. It was constructed from approximately 40 million handmade adobe bricks forming three platforms and four plazas connected by a series of terraces and corridors (figure 2.1). At its base, Huaca de la Luna measures 290 meters north-south and 210 meters east-west. It is located at the foot of Cerro Blanco in the Moche river valley. Some 500 meters to the west is the larger Huaca del Sol, with a dense urban center in between the two platform mounds (Chapdelaine 2001; Uceda 2001). Plaza 3A of Huaca de la Luna is a walled enclosure built around a natural rocky outcrop on the western flank of Cerro Blanco (figures 2.2, 2.3). Plaza 3A is assigned to one of the last stages of construction, ca. AD 450–550, and was excavated from 1995–1997 by Steve Bourget (1997, 1998a, 1998b; Bourget and Millaire 2000). It was found to contain the remains of more than 100 people around the base of the rocky outcrop. The deposit was formed from 15 superimposed layers of human remains in alternating strata of clay and sand. These appear to represent at least five distinct but closely timed events (Bourget 1997, 1998, 2001a, 2001b) associated with a mega–El Niño climatic oscillation event that occurred around AD 550.

3 6   Hamilton

2.2. Plan view of Huaca de la Luna redrawn from Uceda (2001: figure 3). Plazas 3A and 3C are in the southeast sector of the structure.

2.3. An overall view of Plazas 3A and 3C at Huaca de la Luna in 2001 while excavation was ongoing in Plaza 3C. Photo by and courtesy of John Verano.

Ritual Killing, Mutilation, and Dismemberment  37

Plaza 3C is adjacent to and west of Plaza 3A. Plaza 3C was excavated in 1999–2001 by John Verano and Moisés Tufinio (Tufinio 2008; Verano et al. 2007). Here, a markedly different deposit of human remains was found, being temporally and contextually distinct. Plaza 3C contained the bones of at least 33 individuals divided by a floor into two vertically stratified deposits. Radiocarbon and stratigraphic evidence shows that deposition of human remains in Plaza 3C occurred over a few hundred years, perhaps linked to high-status funerary rituals (Verano 2008a) occurring between AD 230–550. The Early Sacrifice Sequence: Plaza 3C The human remains from Plaza 3C individuals were composed of complete, nearly complete, and partial skeletons, 42 sets of grouped remains consisting of articulated elements, clusters of disarticulated bones, and 725 isolated bones. Excavation was still ongoing in 2001 when data collection for this study concluded, and this sample includes approximately two-thirds of the entire assemblage. A count of left femora yielded a minimum number of individuals (MNI) count of 33. The victims were all males, ranging in median age from 14 to 40 years, with an average age of 22 years. The individuals from Plaza 3C exhibited a suite of old, healed perimortem injuries befitting of warriors (see Verano [2005] and Verano and Phillips, chapter 9). In addition, two individuals had ropes knotted around their necks. One individual also had a rope binding his hands behind his back. Rope fragments also were associated with a severed foot. Bleaching and weathering of some bones, along with the presence of numerous empty fly puparia, indicated the bodies had been exposed on the surface and not buried in accordance with “proper” Moche funerary ritual (Millaire 2002). The Later Sacrifice Sequence: Plaza 3A All human remains excavated from Plaza 3A were examined. They consisted of 62 complete and partial skeletons, 30 isolated crania (including nine with articulated mandibles and cervical vertebrae), 211 sets of grouped remains composed of disarticulated bones and articulated elements such as trunks, arms, legs, hands, and feet, and 2,586 isolated bones. A tally of the left talus yielded an MNI of 75, although Backo’s (chapter 3) calculations point to at least 107 individuals present in the

3 8   Hamilton

assemblage. Victims were all male, ranging in median age from 13 to 45 years (average age: 25 years). Verano’s (2001b) interpretation of these men as captives is supported by their sex, robust muscle attachments, partially healed injuries, and old healed cranial and postcranial fractures. Empty muscoid fly puparia and extensive weathering of multiple bones indicate they also underwent a period of exposure and lack of proper burial (Bourget 2001b; Verano 2001a, 2001b). The practice of human sacrifice at Huaca de la Luna clearly spanned several centuries, and the remains of the victims bore unmistakable evidence of a complex range of perimortem traumatic injuries, violent death, and alteration of the bodies (Verano 2001b, 2008a). Data Collection and Analytical Methods Cut marks were examined with an unaided eye, with a 10X hand lens under direct incandescent light, and under a stereomicroscope. They were documented using scale drawings, digital photographs, and color slide film. Cut marks were recorded in terms of number, location on the bone, length in millimeters, orientation relative to one another and to the long axis of the bone, morphology at 20X magnification, and the anatomical structure(s) affected. Assessment of microscopic morphology included data on groove shape, relative width, relative depth, wall morphology, presence of shoulder effects, barbs, and overlapping bone shelves, and damage affecting the cut mark (Binford 1981; Morlan 1984; Shipman 1981; Shipman and Rose 1983). Microscopic features were documented in photomicrographs. Microscopic analysis also included the production of negative molds of the cut marks using polyvinylsiloxane dental impression material. Positive casts were made using a low viscosity epoxy resin mixed with unbleached titanium white medium viscosity acrylic paint to enhance observability. Cross-sections were created by thin-sectioning the positive casts using a trim saw followed by mounting the sectioned casts onto glass slides with an ultraviolet adhesive and placing them into a thinsectioning machine. Positive casts capturing cut mark morphology were examined using an SEM to determine cut mark directionality after Bromage and Boyde (1984). Positive casts in the SEM were examined using an accelerating voltage of 20kV and at magnifications ranging between 15X–2000X, although magnifications between 15–250X were the most

Ritual Killing, Mutilation, and Dismemberment  39

informative. Twenty-four positive casts of well-preserved cervical vertebrae were chosen for directionality analysis using the SEM. A total of 84 cut marks representing a sample of five individuals from above the floor of Plaza 3C and 16 individuals from Plaza 3A were examined in this manner.

Results Cut Mark Patterning in Plaza 3C In Plaza 3C, 12 of the 13 subfloor individuals and 16 of the 17 abovefloor individuals together exhibited cut marks on nearly every skeletal element. Exceptions included the smaller and interior bones of the cranium, the hyoid, the coccyx, two carpals (a trapezoid and pisiform), one tarsal (intermediate cuneiform), and some hand and foot phalanges. Of the 4,456 bones in the Plaza 3C sample, 670 bones exhibited a total of 4,280 cut marks. The most frequently affected elements were the left fibula (22/30 or 73.7%) and left femur (23/33 or 69.6%). Tables 2.1 and 2.2 describe the distribution of cut marks by skeletal element in Plaza 3C as a single sample, as well as within the deposits above and below the floor. Throat-Slitting Fourteen individuals in Plaza 3C exhibited perimortem sharp force trauma consistent with deep laceration of the throat. Superior cervical (C1–C4) vertebrae were more likely to have cut marks than inferior (C5–C7) vertebrae. The C3 (6/23 or 26%) was most frequently cut. Six individuals displayed evidence of cut marks restricted to the anterior aspect of the body and transverse processes of the C1 vertebra, and the anterior aspect of the body of the C2 vertebra, and superior, inferior, and transverse processes of C3–C7 (figure 2.4a). Additional cut marks on the lateral aspect of the C4 vertebra of one individual, and on the anteriolateral aspect of the C6 vertebra of another individual, show cutting strokes across the throat extending to the side of the neck. The skulls and vertebrae of five of the six individuals above were found articulated, indicating their throats were slashed but they were not decapitated. Cut marks consistent with either slashing of the lower throat or removal of the head at the base of the neck were found on the superior aspect of the first rib, as well as in the attachment area for the trapezius muscle of the clavicle and scapula.

4 0  Hamilton Table 2.1. Quantification of Cut Marks on Identifiable Skeletal Elements Deposited in Plaza 3A and Plaza 3C, Huaca de la Luna

Skeletal Element

N Affected* Plaza 3A

% Affected Plaza 3A

N Affected* Plaza 3C

% Affected Plaza 3C 28.0%

Frontal Bone

5/45

11.1%

7/25

Parietal Bone (left)

3/41

7.3%

2/24

8.3%

Parietal Bone (right)

1/42

2.4%

3/24

12.5%

Temporal Bone (left)

0/45

0.0%

3/25

12.0%

Temporal Bone (right)

0/46

0.0%

3/23

13.0%

Occipital Bone

3/52

5.8%

4/27

14.8%

Nasal Bone (left)

0/43

0.0%

4/19

21.1%

Nasal Bone (right)

0/43

0.0%

4/18

22.2%

Malar (left)

0/45

0.0%

2/25

8.0%

Malar (right)

1/45

2.2%

6/25

24.0%

Maxilla (left)

0/44

0.0%

8/25

32.0%

Maxilla (right)

0/44

0.0%

5/25

20.0%

Mandible

0/58

0.0%

10/28

35.7%

C1 (Atlas) Vertebra

3/58

5.2%

1/25

4.0%

C2 (Axis) Vertebra

28/64

43.8%

4/25

16.0%

C3–C7 Vertebrae

15.5%

53/325

16.3%

25/161

Thoracic Vertebrae

0/817

0.0%

70/399

17.5%

Lumbar Vertebrae

0/324

0.0%

32/160

20.0%

First Rib (left)

1/59

1.7%

2/24

8.3%

First Rib (right)

1/57

1.8%

3/25

12.0%

Second Rib (left)

0/50

0.0%

2/16

12.5%

Second Rib (right)

0/50

0.0%

1/15

6.7%

Ribs 3–10 (left)

0/561

0.0%

53/189

28.0%

Ribs 3–10 (right)

4/581

0.7%

40/207

19.3%

Ribs 11–12 (left)

0/72

0.0%

10/28

35.7%

Ribs 11–12 (right)

0/93

0.0%

8/28

28.6%

Sternal Body

0/66

0.0%

1/18

5.6%

Clavicle (left)

2/71

2.8%

13/25

52.0%

Clavicle (right)

0/64

0.0%

5/20

25.0%

Scapula (left)

3/67

4.4%

10/21

47.6%

Scapula (right)

1/64

1.6%

10/21

47.6%

Humerus (left)

2/72

2.8%

10/21

47.6%

Humerus (right)

3/70

4.2%

13/26

50.0%

Radius (left)

1/71

1.4%

8/19

42.1%

Ritual Killing, Mutilation, and Dismemberment  41 Table 2.1. Continued.

Skeletal Element

N Affected* Plaza 3A

% Affected Plaza 3A

N Affected* Plaza 3C

% Affected Plaza 3C

Radius (right)

1/68

1.4%

10/26

38.5%

Ulna (left)

1/74

1.4%

7/22

31.8%

Ulna (right)

3/71

4.2%

7/21

33.3%

Carpals (left)

0/381

0.0%

7/143

4.9%

Carpals (right)

0/379

0.0%

5/131

3.8%

Metacarpals (left)

0/253

0.0%

9/96

9.4%

Metacarpals (right)

3/280

1.0%

10/77

13.0%

Proximal Hand Phalanges (left)

4/178

2.2%

5/76

6.6%

Proximal Hand Phalanges (right)

0/162

0.0%

6/46

13.0%

Middle Hand Phalanges (left)

0/134

0.0%

8/58

13.8%

Middle Hand Phalanges (right)

0/115

0.0%

8/37

21.6%

Os Coxae (left)

0/65

0.0%

16/30

53.3%

Os Coxae (right)

1/62

1.6%

16/31

51.6%

Sacrum

0/72

0.0%

4/31

12.9%

Femur (left)

3/61

4.9%

23/33

69.7%

Femur (right)

3/59

5.1%

19/29

65.5%

Patella (right)

0/57

0.0%

1/24

4.2%

Tibia (left)

0/66

0.0%

15/31

48.4%

Tibia (right)

1/69

1.4%

17/32

53.1%

Fibula (left)

5/70

7.1%

22/30

73.3%

Fibula (right)

1/68

1.5%

18/27

66.7%

Tarsals (left)

0/427

0.0%

15/167

9.0%

Tarsals (right)

0/413

0.0%

9/172

5.2%

Metatarsals (left)

0/294

0.0%

5/111

4.5%

Metatarsals (right)

0/307

0.0%

4/110

3.6%

Proximal Foot Phalanges

3/158

1.9%

1/186

0.5%

Total

151/11352**

1.3%

670/4456**

15.0%

* The number of elements with cut marks over the number of identifiable elements present. ** Element fragments of indeterminate side with cut marks are not included in this table.

4 2   Hamilton Table 2.2. Quantification of Cut Marks on Identifiable Skeletal Elements Deposited Below and Above the Floor of Plaza 3C, Huaca de la Luna Skeletal Element Frontal Bone

N Affected* % Affected N Affected* % Affected Below the Floor Below the Floor Above the Floor Above the Floor 4/9

44.4%

3/16

18.8%

Parietal Bone (left)

1/10

10.0%

1/14

7.1%

Parietal Bone (right)

1/10

10.0%

2/14

14.3% 6.7%

Temporal Bone (left)

2/10

20.0%

1/15

Temporal Bone (right)

2/9

22.2%

1/14

7.1%

Occipital Bone

2/10

20.0%

2/17

11.8%

Nasal Bone (left)

3/6

50.0%

1/13

7.7%

Nasal Bone (right)

3/6

50.0%

1/12

8.3%

Malar (left)

1/10

10.0%

1/15

6.7%

Malar (right)

3/10

30.0%

3/15

20.0%

Maxilla (left)

5/10

50.0%

3/15

20.0%

Maxilla (right)

3/9

33.3%

2/16

12.5%

Mandible

7/13

53.8%

3/15

20.0%

C1 Atlas Vertebra

1/11

9.1%

0/14

0.0%

C2 Axis Vertebra

1/10

10.0%

3/15

20.0%

C3–C7 Vertebrae

10/70

14.3%

15/91

16.5%

Thoracic Vertebrae

56/178

31.5%

14/221

6.3%

Lumbar Vertebrae

22/80

27.5%

10/80

12.5%

First Rib (left)

1/5

20.0%

1/19

5.3%

First Rib (right)

2/9

22.2%

1/16

6.3%

Second Rib (left)

0/2

0.0%

2/14

14.3%

0/3

0.0%

1/12

8.3%

Ribs 3–10 (left)

Second Rib (right)

10/52

19.2%

43/137

31.4%

Ribs 3–10 (right)

15/67

22.4%

25/140

17.9%

Ribs 11–12 (left)

0/4

0.0%

10/24

41.7%

Ribs 11-12 (right)

0/2

0.0%

8/26

30.8%

Sternal Body

0/1

0.0%

1/17

5.9%

Clavicle (left)

5/7

71.4%

8/18

44.4%

Clavicle (right)

1/4

25.0%

4/16

25.0%

Scapula (left)

3/7

42.9%

7/14

50.0%

Scapula (right)

2/5

40.0%

8/16

50.0%

Humerus (left)

4/6

66.7%

6/15

40.0%

Humerus (right)

6/10

60.0%

7/16

43.8%

3/5

60.0%

5/14

35.7%

Radius (left)

Ritual Killing, Mutilation, and Dismemberment  43 Table 2.2. Continued.

Skeletal Element

N Affected* Below the Floor

% Affected N Affected* % Affected Below the Floor Above the Floor Above the Floor

Radius (right)

4/8

50.0%

6/18

33.3%

Ulna (left)

4/8

Ulna (right)

1/5

50.0%

3/14

21.4%

20.0%

6/16

37.5%

Carpals (left)

0/18

0.0%

7/125

5.6%

Carpals (right)

2/30

6.7%

3/101

3.0%

Metacarpals (left)

0/16

0.0%

9/80

11.3%

Metacarpals (right)

4/21

19.0%

6/56

10.7%

Proximal Hand Phalanges (left)

0/13

0.0%

5/63

7.9%

Proximal Hand Phalanges (right)

0/5

0.0%

6/41

14.6%

Middle Hand Phalanges (left)

0/9

0.0%

8/49

16.3%

Middle Hand Phalanges (right)

0/6

0.0%

8/31

25.8%

Os Coxae (left)

12/15

80.0%

4/15

26.7%

Os Coxae (right)

12/15

80.0%

4/16

25.0%

Sacrum

2/16

12.5%

2/15

13.3%

Femur (left)

12/16

75.0%

11/17

64.7%

Femur (right)

9/12

75.0%

10/17

58.8%

Patella (right)

1/11

9.1%

0/13

0.0%

Tibia (left)

7/13

53.8%

8/18

44.4%

Tibia (right)

9/14

64.3%

8/18

44.4%

Fibula (left)

12/16

75.0%

10/14

71.4%

Fibula (right)

7/11

63.6%

11/16

68.8%

Tarsals (left)

7/55

12.7%

8/112

7.1%

Tarsals (right)

5/62

8.1%

4/110

3.6%

Metatarsals (left)

3/36

8.3%

2/75

2.7%

Metatarsals (right)

0/40

0.0%

4/70

5.7%

Proximal Foot Phalanges

1/67

1.5%

0/119

0.0%

308/1542**

20.0%

362/2914**

12.4%

Total

*The number of elements with cut marks over the number of identifiable elements present. **Element fragments of indeterminate side with cut marks are not included in this table.

4 4   Hamilton

A

b

2.4. Examples of perimortem treatment among the sacrifice victims in Huaca de la Luna Plaza 3C consistent with (A) extensive throat-slitting and (B) decapitation. Photos by Laurel Hamilton.

Decapitation Cut marks consistent with decapitation were observed on the lateral or posterior aspects of the superior articular facets, transverse processes, lamina, inferior articular facet, and spinous processes of C2–C7 vertebrae (figure 2.4b) in four individuals; each was missing their skull and at least the atlas vertebrae. The four also had displaced cervical vertebrae and anterior cut marks, indicating a combination of throat-slitting and decapitation. Defleshing Defleshing of the face (figures 2.5a, 2.5b) was indicated by cut marks in the glabellar region, superciliary arches, and squamous portion of the frontal bone. Other cuts were found along the temporal bone, the left and right parietal bones, and the squamous portion of the occipital bone. Postcranially, cuts were found on the spinous processes and posterior aspects of vertebral transverse processes, the laminae of the thoracic and lumbar vertebrae, the lateral aspect of the pedicle and anterior and lateral aspects of the transverse processes of the lumbar vertebrae, and the superior and external aspects of ribs 2–12 (figure 2.6a). Other affected locations included the anterior aspect of the body of the sternum, the diaphysis of the clavicle, the acromion process, subscapular

Ritual Killing, Mutilation, and Dismemberment  45

fossa, axillary border, superspinous fossa, spine and infraspinous process of the scapula, the diaphysis and medial and lateral epicondyles of the humerus, the head, neck, and diaphysis of the radius, and the diaphysis and the anterior and medial aspects of the coronoid process of the ulna (figure 2.6b). Defleshing of the hand was also suggested by cuts on the dorsal and palmar aspects of metacarpals and proximal hand phalanges. Excarnation of the lower body was inferred from cut marks on the body of the A

b

2.5. Examples of perimortem treatment among the sacrifice victims in Huaca de la Luna Plaza 3C, demonstrating (A) mutilation of the noseand (B) removal of the lips. Figure 2.5a by John Verano. Figure 2.5b by Laurel Hamilton

A 2.6. Examples of postmortem treatment documented in Plaza 3C involving (A) defleshing of the thorax and defleshing of upper limbs (B) and (C) lower limbs. Photos by Laurel Hamilton

b

c

4 6   Hamilton

pubis, the lateral aspect of the iliac crest, iliac blade, iliopubic ramus and ischiopubic ramus, the medial aspect of the os coxae, the greater and lesser trochanters and diaphysis of the femur, the diaphysis of the tibia, the head, neck, and diaphysis of the fibula, and all aspects of the tarsal bones. Cuts on the calcaneal tuberosity of the left calcaneus of one individual and on the diaphyses of four metatarsals of four other individuals may also represent defleshing of the foot. Dismemberment Disarticulation of the thorax of one Plaza 3C victim is suggested by cut marks on the anterior and inferior aspects of the superior articular facets, transverse processes, and laminae of the thoracic vertebrae. The removal of the arm at the shoulder joint was indicated by cuts on the diaphysis of the clavicle and the subscapular fossa, axillary border, spine, and infraspinous fossa of the scapula. Also, the associated humerus was missing. In two other cases, only detached—but articulated—arms were present. For other individuals, removal of the arm at the shoulder was indicated by cut marks penetrating joint capsules (i.e., on the surgical neck and proximal diaphysis of the humerus). In all but one case, the humerus was not in anatomical position and the associated clavicle and scapula were missing. Disarticulation of the forearm was indicated by cuts on the distal diaphysis of the humerus together with the absence of one or both forearm bones. Cut marks on the styloid processes of the left and right radii of one individual most likely represent the removal of the hands at the wrist, and both hands were missing. Detachment of the hand at the wrist also is suggested by cut marks on the carpals of extra hands found with one individual and on the carpals of the isolated hand of another individual. Cuts on the distal diaphyses and distal articular surfaces of the left and right first proximal phalanges and all cut marks on the middle phalanges of two individuals indicated removal of the thumb and fingers at the interphalangeal joints (figure 2.7a). The absence of all distal phalanges of both individuals further supports this interpretation. In the lower body, cut marks in the areas of muscle and ligament attachment on both the medial and lateral aspects of the os coxae may be associated with the removal of the lower limb at the hip joint. In one case, cut marks were observed around the rim of the acetabulum of the right os coxae as well as on the neck, greater trochanter, and proximal fourth of the diaphysis of the right femur of one individual, and on

Ritual Killing, Mutilation, and Dismemberment  47

A

b

2.7. Other examples of postmortem treatment of Plaza 3C included (A) dismemberment (in this case, of fingers) and (B) partial dismemberment of the leg from the hip joint. Photos by Laurel Hamilton.

the neck and proximal diaphysis of the left femur of another individual. However, in both cases, the affected femur was found in anatomical articulation with the associated os coxae (figure 2.7b). If dismemberment was the final objective, this goal was not achieved, despite numerous cut marks. Disarticulation of the pelvis may explain cuts in the area

4 8   Hamilton

of attachment of the external oblique muscles on the os coxae of three individuals, supported by the observation that their os coxae were found disarticulated. Cut marks on the lateral aspect of the iliopubic ramus of three individuals may represent mutilation or removal of the genitals. One individual had cuts on the distal diaphysis of the right femur, a cut mark just below the fibular facet on the posterior aspect of the left tibia, and cuts on the neck and posterior aspect of the proximal diaphysis of the left fibula. The absence of the right patella—and the apparent subdiaphragmatic insertion of the right tibia and fibula and elements of the lower left leg and foot into this man’s chest cavity—indicate disarticulation of the legs at the knees. For another individual, disarticulation at the knee was demonstrated by cut marks on the posterior aspect of the patella in conjunction with a horizontal section of bone that was removed in the perimortem interval. Manual dislocation of the hyperflexed knee joints in one individual explains cut marks on the proximal diaphyses of the left and right tibiae, styloid process, neck and proximal diaphysis of the left fibula, and proximal diaphysis of the right fibula. Detachment of the foot at the ankle was indicated by cut marks on the distal diaphysis and medial malleolus of the tibia, the distal diaphysis or lateral malleolus of the fibula, and on the tarsal bones. In all cases, the affected foot or feet were missing, found in isolation, or were disarticulated. Cut marks together with the absence of associated foot phalanges suggest the removal of toes. Cuts on the proximal diaphysis of an isolated first foot phalanx may represent dismemberment. Possible Chest-Opening An argument can be made for the prying open of the chest cavity (and possible removal of the heart) of two individuals in Plaza 3C. With one individual, all cut marks on the ribs were located on the internal surface of the bone and traversed the vertebral portion of the shaft, neck, or tubercle of the bone. Furthermore, the ribs of both individuals were represented by only fragments and exhibited perimortem fractures of the neck or shaft just anterior of the tubercle. Cut Mark Patterning in Plaza 3A Of the 11,352 bones from Plaza 3A, 151 bones exhibited a total of 777 cut marks (table 2.1). Cut marks predominantly affected cervical vertebrae but also were observed on the skull, ribs, left clavicle, scapulae, humeri,

Ritual Killing, Mutilation, and Dismemberment  49

radii, ulnae, metacarpals, hand phalanges, right os coxae, femora, tibiae, fibulae, and foot phalanges. Throat-Slitting Sharp force trauma was found on the C1–C7 vertebrae of 61 individuals in Plaza 3A. Superior (C1–C4) vertebrae were more likely to have cut marks than inferior (C5–C7) vertebrae, with the C3 (32/59 or 54.2%) most commonly affected. Of these, 45 individuals had their throats slit—as indicated by cuts restricted to the anterior aspect of the body, dens, and anterior transverse processes of the C2, the anterior surface of the body, superior articular processes, and transverse processes of other cervical vertebrae (figure 2.8a). Decapitation Nine individuals in Plaza 3A were decapitated, as indicated by cut marks on the inferior aspect of the anterior arch, the anterior aspect of the right transverse process, the anterior margin of the C1, and the right inferior articular facet of the C2. Further markers of decapitation involved absence of a skull, the absence of cervical vertebrae superior to those with cut marks, the lack of postcrania inferior to the cervical vertebrae, and cervical vertebrae missing portions of bone perimortem, especially the left and right lateral margins of the superior aspect of the vertebral body owing to the complete passage of the knife through the bone. Seven individuals had their throats slit followed by decapitation as indicated by cervical vertebrae with cut marks characteristic of both activities. Cut marks consistent with slashing of the lower throat or removal of the head at the base of the neck were found on the superior aspect of the first rib and on the superior and posterior left clavicles. Defleshing Cut marks consistent with laceration of the scalp, including flaying of the forehead, were noted on the squamous portion and left superciliary arch of the frontal bone, right and left parietal bones, and squamous portion of the occipital bone. A few other cases involved cut marks on the external aspect of ribs, the spine, superior border, infraspinous fossa and inferior angle of the scapula, lateral epicondyle of the humerus, diaphysis of the humerus, radius, ulna, hand phalanges, femur, tibia, and fibula, and the iliac blade and ischial tuberosity of the os coxae.

5 0  Hamilton

Dismemberment Mutilation of the fingers and toes is suggested by cut marks on the diaphyses of metacarpals and hand and foot phalanges (Figure 2.8b), but removal of fingers is indicated by cut marks on or near articular facets of the hand phalanges and the fact that all middle and distal hand phalanges were absent in each case.

A

b

2.8. Examples of perimortem treatment among Huaca de la Luna Plaza 3A sacrificial victims, with sharp force trauma produced from (A) throat-slitting (Cráneo XVIIIa from Plaza 3A) and (B) mutilation of fingers. Photos by Laurel Hamilton.

Ritual Killing, Mutilation, and Dismemberment  51

Cut Mark Morphology and Directionality Ninety-three percent (3981/4280) of the Plaza 3C and 95% (740/777) of the Plaza 3A cut marks both in cross-section and wall morphology possessed an identical morphological pattern: V-shaped grooves with straight, sheer walls (figures 2.9, 2.10, 2.11). Although a few cuts had relatively coarse walls or one straight wall and one coarse wall, in nearly all cases these variations could be attributed to damage, weathering, or bone structure. Only 1.4% of Plaza 3A cut marks and 3% of Plaza 3C cut marks were U-shaped. In nearly all cases, the U-shaped crosssection could be attributed to shallow depth or multiple cuts sharing the same groove. Shoulder effects or barbs were absent. Eleven percent (88/777) of the Plaza 3A cut marks and 6.7% (287/4280) of the Plaza 3C cut marks exhibited overlapping bone shelves (figure 2.12). The only exceptions were relatively wide, deep, V-shaped chop marks observed on an occipital squama, the proximal diaphysis of a humerus, and the midshaft of a femur of three different individuals from Plaza 3A. Of the 24 cervical vertebrae cast in resin, 98% (82/84) of cuts were

2.9. Bone smear (boxed area) on the floor of a cut mark on the anterior aspect of the C3 vertebra of HG99-5 from Plaza 3C. Directionality of cutting went from left to right. Image by Laurel Hamilton.

5 2   Hamilton

2.10. Scanning electron micrograph illustrating the straight, sheer walls of a cut mark on the left twelfth rib of HG99-3 from Plaza 3C. The black arrow denotes the superior margin of the cut mark, and the white arrow indicates the main groove of the cut mark. The inferior wall of the cut mark is not clearly discernible due to a 30° angle of tilt. Image by Laurel Hamilton.

A

b

2.11. Representative cross-sectional profile views of cut marks on human bones from (A) Plaza 3A and (B) Plaza 3C, Huaca de la Luna. Photos by Laurel Hamilton.

Ritual Killing, Mutilation, and Dismemberment  53

2.12. Two parallel cut marks on the spine of the left scapula of HG96-44 from Plaza 3A, Huaca de la Luna. Note the presence of an overlapping bone shelf consistent with use of a metal knife. Photos by Laurel Hamilton.

located on the anterior aspect of the vertebral body or right transverse process (table 2.3). Oblique faults and bone smears were observed. Oblique chipping was absent. Typically, only one criterion was present in each cut. However, in three cases—one from Plaza 3C and two from Plaza 3A—both bone smears and oblique faults were observed in the same cut. Approximately 58% (49/84) of the cut marks lacked evidence of directionality. As a whole, more individuals had their throats cut from left to right than those cut from right to left or from both directions (table 2.4). Within the Plaza 3C sample, three individuals had their throats slit from both directions. Ten individuals from Plaza 3A exhibited left to right throat-slitting, while only one individual in each category had his throat slit from right to left or from both directions. In the remaining four cases from Plaza 3A, directionality could not be determined.

5 4   Hamilton Table 2.3. Counts of Directionality Criteria and Cutting Strokes in the Plaza 3A and 3C Samples Sample

Cut Marksa Bone Smearsb Oblique Faultsc Noned Left to Righte

Right to Leftf

Plaza 3A

55

16

9

32

16

8

Plaza 3C

29

11

2

17

7

6

Total

84

27

11

49

23

14

Number of cut marks observed Number of bone smears observed c Number of oblique faults observed d Number of cut marks for which no directionality criteria were observed e Number of left to right cutting strokes f Number of right to left cutting strokes a b

Table 2.4. Throat-Slitting Directionality Results by Individual Sample

Individualsa

Left to Rightb

Right to Leftc Both Directionsd

Direction N/Ae

Plaza 3A

16

10

1

1

4

Plaza 3C

5

1

1

3

0

Total

21

11

2

4

4

Number of individuals represented per data sample with cut marks on the anterior aspect of cervical vertebrae indicative of throat-slitting b Number of individuals whose throats were cut from left to right c Number of individuals whose throats were cut from right to left d Number of individuals whose throats were cut from both left to right and right to left e Number of individuals for whom directionality of cuts to throat is not available a

Discussion Victim Treatment The skeletal assemblage of Plaza 3C epitomizes the extremes and variations of Moche sacrifice victim treatment, particularly in terms of comparisons between the subfloor and above-floor deposits. As noted in table 2.2, cut marks are present on more skeletal elements above the floor than below the floor, especially on ribs and hand elements. On bones deposited below the floor, cut marks occur most frequently on the left and right os coxae, whereas above the floor the most affected element is the right fibula. Most Plaza 3C skeletons have cuts on diverse

Ritual Killing, Mutilation, and Dismemberment  55

bones spanning multiple anatomical regions (figure 2.13). The number of cuts per individual is highly variable, ranging from 28 to 343 for individuals in the subfloor deposit and from 1 to 288 for individuals in the above-floor deposit. As a group, the subfloor individuals received more complex treatment than the above-floor individuals; there are more cut marks overall and indications of facial mutilation, while evidence of excision of both lips (as opposed to the upper lip only) is limited to above-floor remains. Inferred genital mutilation and disarticulation of the pelvis is found exclusively below the floor, while mutilation of the hands and feet is restricted to above the floor. The primary objective at Plaza 3C was to slash throats and deflesh bodies, especially arms and legs. A primary focus on defleshing (not dismemberment) is further indicated by the fact that many of the skeletons were largely anatomically articulated in situ. Smaller subsets of cuts were

2.13. Cut mark locations on the individual designated HG99-5 from Plaza 3C, demonstrating the common pattern of shaded multiple cut marks spanning multiple anatomical regions. Missing elements shaded in black. Drawing by Laurel Hamilton.

5 6   Hamilton

observed on or near joint surfaces. Dismemberment was highly selective in each case but, paradoxically, did not involve repetitive selection of body parts for removal. Cut marks also reflect laceration of the scalp, mutilation around the eyes, nose, cheeks, and lips, mutilation of the hands and feet, and possible mutilation of the genitals. Cuts exhibited by seven individuals at the base of the body, rami, and inferior aspect of the mandible may reflect excision of the tongue. If endured during the premortem interval, these injuries would reflect particularly violent torture preceding death. Such mistreatment may have held a ritualistic performative aspect. Scalp and tongue lacerations bleed profusely, and multiple small cuts near the eye would have produced considerable pain and bleeding. One can speculate that both these acts and the resulting bloodied appearance of the victim would have had a terrorizing and intimidating effect on the victims and other supplicants or attendees. Perhaps these actions involved the transformation of a captive into a sacrifice victim. The extent to which this process could progress is seen in the case of Plaza 3C Entierro 3 with sharp force trauma reflecting throat-slitting, decapitation, defleshing of the trunk, right arm, and both legs, disarticulation of the thorax, and removal of the right arm and both feet. Manners of Killing: Tool Types and the Physicality of Sacrifice All cut marks in this study appear to have been made with metal tools. Further, on nearly every bone where U-shaped cuts were found, V-shaped cuts were also present; observed morphological variation appears related to differences in the amount of force and applied pressure or varying sharpness of the tool edge. The V-shaped chop marks observed in three individuals from Plaza 3A were associated with impact scars and probably represent perimortem blows with a sharp-bladed weapon. Metal tools are also consistent with the presence of dozens of cuts with overlapping bone shelves. Injuries associated with throat-slitting and decapitation were probably made with a tumi or chisel-edged knife (as documented in Sipán Tombs 1 and 3, for example). Although the Moche did not use obsidian tools (Glenn Russell, personal communication 1997), they did utilize shale, diorite, rhyolite, andesite, and basalt (Claude Chapdelaine, personal communication 1999; Shimada 1994a). Still, no stone tumis have ever been documented, and it seems clear that the Moche consistently—and consciously—used metal implements in rituals of human sacrifice.

Ritual Killing, Mutilation, and Dismemberment  57 Table 2.5. Assailant-Victim Positioning Possibilities During Throat-Slitting Events Based on Directionality of Cut Marks Located on the Anterior Aspect of Cervical Vertebrae A) Cut Mark Direction: Left to Right Handedness of Assailant

Position of Assailant Relative to Victim

Movement of Assailant’s Arm When Slitting Victim’s Throat

Right

Facing Victim

Extension

Right

Behind Victim

Flexion

Left

Facing Victim

Flexion

Left

Behind Victim

Extension

B) Cut Mark Direction: Right to Left Handedness of Assailant

Position of Assailant Relative to Victim

Movement of Assailant’s Arm When Slitting Victim’s Throat

Right

Facing Victim

Flexion

Right

Behind Victim

Extension

Left

Facing Victim

Extension

Left

Behind Victim

Flexion

The directionality study reconstructed elements of physicality and positioning of the assailant relative to cases of throat-slitting (table 2.5). Cross-cultural research among modern humans shows approximately 90% of people are right-handed (Glassman and Dana 1992). Two possibilities emerge. First, a right-handed assailant stood in front of the victim to cut his throat from left to right. Face-to-face positioning would mean that cutting-stroke motions would have to push across the throat in a counterintuitive and awkward fashion. The assailant would also be rapidly drenched in blood spray. The second, more likely scenario involves a right-handed assailant behind the victim; with the greater leverage and mechanical advantage this positioning conferred, the victim’s throat was cut with a tool held in the assailant’s right hand. Where cuts traverse cervical vertebrae from right to left, only one scenario is likely with a right-handed sacrificer: the assailant stood in front of the victim. A right-handed assailant positioned behind the victim would again require a counterintuitive and potentially difficult feat.

5 8   Hamilton

In cases where the cut marks advance from both left to right and right to left, it is most likely that an assailant stood behind the seated victim and cut the throat using a sawing motion. The assailant, using the left hand, was probably capable of capturing the victim’s blood into a goblet or similar container. However, these actions probably entailed the coordinated efforts of at least one additional individual or assistant, who presumably restrained the victim, pulled the head back by the hair to hyperextend the neck, and collected the blood. Plaza 3C Versus Plaza 3A: Change Over Time? Various commonalities link the activities in the earlier Plaza 3C and later Plaza 3A, including victim age, sex, and a prior history of violent encounters. Throat-slitting was also equally frequent. However, the similarities end there. While most individuals from Plaza 3A simply had their

2.14. Cut mark locations on the individual designated HG96-1 from Plaza 3A, the only victim in Plaza 3A to demonstrate cut marks akin to the victims in Plaza 3C. Missing elements shaded in black. Drawing by Laurel Hamilton.

Ritual Killing, Mutilation, and Dismemberment  59

2.15. Comparison of the types of perimortem and postmortem treatment suggested by cut marks on individuals represented by articulated and semi-articulated skeletons from Plaza 3C (n =3 0) and Plaza 3A (n = 62), Huaca de la Luna. Note that facial mutilation is not represented on the graph for Plaza 3A because the only cut marks on facial bones were present on the right malar of an isolated cranium. Image by Laurel Hamilton.

throats slit, the treatment of the Plaza 3C victims as noted earlier was far more complex. Fifteen percent of the Plaza 3C bones had cut marks, as opposed to only 1.3% of the bones from Plaza 3A. Among Plaza 3C remains, the most frequently cut bone was the left fibula, whereas within the Plaza 3A sample, cut marks occurred most frequently on the C2 vertebra. Nearly 95% of the victims represented by articulated and semi-articulated skeletons from Plaza 3C had cut marks, compared to 44% of the skeletons from Plaza 3A. Only one skeleton from Plaza 3A, HG96-1, had multiple cut marks on the midshafts of upper and lower long bones, mirroring patterns seen in Plaza 3C (figure 2.14). Figure 2.15 compares the types of perimortem and postmortem treatment suggested by cut marks on individuals, represented by articulated and semi-articulated skeletons from Plaza 3C and Plaza 3A. Indeed, most activities represented by cut marks on Plaza 3C bones are also represented in Plaza 3A, including throat-slitting, decapitation, scalp and facial lacerations, and mutilation of the hands and feet—but they occur in far lower frequencies in the later Plaza 3A. Bundles of articulated

6 0  Hamilton

arms and legs, such as Entierro 6 in Plaza 3C, were not present in Plaza 3A. There is very little evidence of sharp force–related dismemberment in Plaza 3A (Backo, chapter 3). Clearly, there are diachronic differences between these two assemblages. However, is change over time really the right question to pose? Dissimilarities in treatment between Plaza 3C and Plaza 3A could signify more about different motivations and unique circumstances under which the sacrifices took place. Based on stratigraphic and radiometric data, Plaza 3C represents a gradual accumulation of victims over a few hundred years. Verano (2008a:207) states “the sub-floor remains were incorporated in the sand and adobe fill that was used to raise the level of the plaza during its construction,” while “the above-floor remains . . . were [either] covered with adobe fragments or simply left to be buried by accumulating windblown sand.” The extensive postmortem processing strongly suggests that Plaza 3C was a repository for remains of victims sacrificed elsewhere. In contrast, Plaza 3A appears as a primary sacrificial site where the bodies were cast around the rocky outcrop following death (Verano 2008a). Bourget (1998b, 2001a) argues that the remains deposited in Plaza 3A were directly associated with the documented mega–El Niño climate event that affected the north coast ca. AD 550 (Shimada et al. 1991). Several skeletons were found in alternating layers of hardened clay and windblown sand, indicating sacrifices occurred both around and after episodes of torrential rainfall. Imprints of fleshed arms and legs around the bones of some skeletons in Plaza 3A showed some bodies were deposited while the mud was still fresh (Steve Bourget, personal communication 2013). Although thin layers of silt left by episodes of light rainfall were present in Plaza 3C, none of the skeletons were associated with these layers; Plaza 3C is not associated with El Niño phenomena (Verano 2008a). Ritual killing in the two plazas truly seems situated in distinct historical settings and ritual spaces: Plaza 3C is associated with liturgical and high-status funerary rites over at least a few centuries (Verano 2008a), while Plaza 3A appears to have involved rituals associated with an intense El Niño event. Conjunctions and Disjunctions Between Physical and Iconographic Evidence While many parallels between physical and iconographic evidence of Moche sacrifice exist, a one-to-one correspondence appears elusive. One

Ritual Killing, Mutilation, and Dismemberment  6 1

of the clearest matches involves the artistic depiction of throat-slitting in the Sacrifice Ceremony (Donnan and McClelland 1999). Abundant physical evidence demonstrates throat-slitting was a standard killing technique at Huaca de la Luna. Also, the Moche represented dismembered sacrifice victims by their severed heads, arms, and legs, with ropes tied around them to facilitate transport and display (Donnan 1978). Such representations are consistent with cut marks on or around the joint surfaces of bones, missing body parts, and, in Plaza 3C, rope fragments found around the wrists and ankles of isolated arms and legs (Verano et al. 2007). In Moche sacrifice scenes, the assailant was always shown standing, often facing the victim (Donnan 2004; Donnan and McClelland 1999) or standing to the victim’s right side (Donnan 1978). Right-handed and left-handed assailants were portrayed more or less equally. The hand not clutching a tumi was almost always holding the victim’s hair or a goblet (Donnan 2004). The victim is typically portrayed seated but in some cases is standing (Donnan and McClelland 1999). The scenarios suggested earlier to explain the directionality of cuts on cervical vertebrae bear many similarities to the iconography. Still, the physical evidence suggests degrees of artistic license in the idealized positioning of the assailant or victim. In Moche art, some prisoners are depicted with mutilated noses and lips, while others exhibit mutilated genitals or legs (Gálvez and Briceño 2001:140; Larco Hoyle 2001b: fig. 48). This is consistent with the eight individuals from Plaza 3C who probably endured perimortem nose or lip removal. One of those individuals, in addition to five others, had cuts on the anterior os coxae suggesting mutilation or removal of the genitalia. Other forms of documented mistreatment, including scalp lacerations, wounds around the eyes, stripping of flesh, and mutilation of the hands, legs, and feet, are absent in Moche art. Representations of animated skeletons could represent the end result of the defleshing process. Skeletons played a very important role in Moche culture and are shown in a variety of activities, including playing musical instruments, dancing, and engaging in sexual acts with living women (Bourget 2006; Donnan and McClelland 1999) involving metaphors connecting fertility and the life-generating qualities of dead beings. Moche art does not appear to ever depict chest mutilation. Zighelboim (1995) interprets the cone-shaped objects held in the claws of the two-headed serpent that supports the upper register of the Sacrifice

6 2   Hamilton

Ceremony as hearts extracted from sacrificial victims. In addition, he claims the scene in the upper-right corner of figure 7.5 in Donnan (2004) also represents heart extraction. While Zighelboim’s (1995) identification of the objects in the two-headed serpent’s claws as hearts is possible, the latter scene depicts throat-slitting—not heart extraction. Evidence for Moche practices of prying open the chest is exclusive to the archaeological record.

Conclusion Macro- and microscopic analysis of cut mark location and morphology made it possible to identify a diverse range of mutilation and dismemberment at Huaca de la Luna—and provided new data on the tools and techniques associated with Moche sacrifice and relationships between physical trauma and iconographic representations of ritual killing. On some levels, Moche sacrifice was systematically consistent in the cutting of most of their victim’s throats. But in the rituals that preceded the Plaza 3A killings toward the end of Moche IV, their earlier activities possessed many highly variable elements involving extensive mutilation, defleshing, possible torture, and dismemberment. Second, the diachronic perspective provided in comparisons between Plazas 3C and 3A suggests that their differences may be less involved with changes over time rather than the more immediate contexts and circumstances motivating these killings. The two assemblages reflect an ideological system that was both highly ritualized and brutally efficient in times of crisis. Third, cut mark morphology indicates that metal tools were used to commit these acts, as the sacrificers assumed positions in front of and behind their victims. In addition, whereas the efficacy of a stone tool would be limited, a metal tool could be reused many times and the tool itself able to retain some of the power imbued by the ritual acts. Fourth, while general parallels do exist between Moche art and the physical evidence of sacrifice, some types of treatment were clearly not depicted, but the reason why is currently unclear. Perhaps particular treatments had greater significance and thus were deemed more worthy of illustration than others. In any case, Moche artistic portrayals of ritual killing were not fully representational and cannot be interpreted literally as an exact script. Future cut mark analyses will be key to advancing the study of Andean sacrifice. Future analysis of the remaining one-third of the human bones excavated from Plaza 3C, along with the new finds of Late Moche killings

Ritual Killing, Mutilation, and Dismemberment  6 3

from Huaca de la Luna Platform III and sacrificial victims from El Brujo and Dos Cabezas, will further advance the study of Moche sacrifice. Larger datasets will be amenable to various statistical analyses of patterning to expand these initial directionality studies. Comparison with cut mark patterning on faunal remains from Moche ritual and domestic contexts will permit further evaluation of differential treatment of sacrificed humans versus butchered animals. Elemental analysis using a micro x-ray fluorescence spectrometer of human and faunal bones with deep, debris-filled cut marks may provide yet more definitive assessments of tool types, regardless of the activity or individuals involved. These advances will ultimately allow us to build more detailed and confident understanding of ritual violence in the Moche world and beyond. Acknowledgments The Fulbright-Hays Doctoral Dissertation Research Abroad Program (Grant #P022A000057), the National Science Foundation (Doctoral Dissertation Improvement Grant #0075174), the Middle American Research Institute at Tulane University, and the Tulane University Department of Anthropology provided financial support for this research. Additional generous support and assistance was provided by John Verano, Moisés Tufinio, Santiago Uceda, Steve Bourget, César Gálvez, Christopher Donnan, Alana Cordy-Collins, Rose Tyson, Tom Wake, Jean Hudson, Kendall Campbell, George (Wolf ) Gumerman IV, Glenn Russell, Ventura Perez, Robert Dotson, Pierre Burnside, Ronald Parsley, T. R. Kidder, Tony Ortmann, Lori Roe, Jeb Card, Kit Nelson, Margaret Clarke, Harvey Bricker, Trenton Holliday, Conrad Hamilton, Gary and Janice Anderson, Florencia Bracamonte, Ulla Holmquist Pachas, Luis Jaime Castillo, Tania Delabarde, Juan Julio Bracamonte, the Montoya Vera family, and the Patsías Valle family. Haagen Klaus, J. Marla Toyne, and three anonymous reviewers provided editorial comments on previous drafts of this chapter.

6 4   Backo

chapter three

The Taphonomy of Ritual Killing on the North Coast of Peru Perspectives from Huaca de la Luna and Pacatnamú Heather C. Backo Joint POW/MIA Accounting Command Central Identification Laboratory, Oahu, Hawaii

Ritual killing was a core element of belief systems in many cultures on the north coast of ancient Peru—today known through decades of research on iconography (Alva and Donnan 1993; Cordy-Collins 1996a, 1996b, 2001a; Donnan 2004; Hudson Museum 1998; UbbelohdeDoering 1952), archaeological discoveries (Bourget 1997, 1998a, 1998b; Cordy-Collins 2001a, 2011b; Tufinio 2008; Tufinio et al. 2009; Verano 1986a, 1998, 2000, 2001a, 2001b, 2005; Verano et al. 2008; Verano and Backo in press), and the physical anthropology of sacrifice victims themselves (Verano 1998, 2000, 2001a, 2001b, 2007). The Moche (AD 100–850) appear to be the first society on the north coast of Peru to practice large-scale ritual killing and were certainly not the last (Klaus et al. 2010; Toyne 2011a, 2011b; Verano 1986a; Wester et al. 2010a). Questions still remain as to the nature of these rituals, especially the broader sequence of events that followed acts of sacrifice as well as the ways in which the living interacted with the bodies of the sacrificial dead. Sacrificial assemblages on the north coast of Peru tend to be complex depositional contexts, with multiple layers of disarticulated, commingled, and scattered human remains that appear fairly catastrophic or disorganized when first examined (Bourget 1997, 1998b). Yet victims were regularly left exposed on the surface, remaining within the world of the living. In this chapter, a taphonomic approach is employed to

The Taphonomy of Ritual Killing  6 5

characterize the nature and sequence of postmortem circumstances of sacrificial victims. In a comparative case study–based examination of burial taphonomy in two temporally and culturally distinct settings (Huaca de la Luna Plaza 3A and the Pacatnamú mass grave), this work focuses on four primary questions within and between these two contexts: (1) were there regular patterns in the distribution, orientation, and articulations of the recovered human remains; (2) were specific skeletal elements represented at significantly lower rates than others; (3) are patterns of disarticulation and missing elements attributable to human activity or natural taphonomic processes; and (4) what do taphonomic perspectives reveal about the interactions between the living and their sacrificial dead vis-à-vis the meanings, motivations, and ethos underscoring ritual killing? This chapter also introduces a new configuration of methods to characterize mortuary contexts with complicated postmortem histories.

Background: Taphonomy, Commingling, and Physical Complexity in Mass Burial Sites Taphonomy, as introduced by Efremov (1940), was first conceived as a subfield of paleontology focused on transition of remains from the biosphere to the lithosphere. Since then, taphonomy has been defined as “the study of fossils in their geological contexts with a view to sorting out the factors intervening between death and definitive burial that . . . render paleoecological reconstruction difficult” (Dodson 1980:631), and as “the study of processes of preservation and how they affect information in the fossil record” (Behrensmeyer et al. 2000:103). Over the last several decades, taphonomy has evolved beyond its paleontological roots to contribute substantially to anthropological science, advancing understandings of past and present human behavior (e.g., Duday 2006, 2009; Haglund and Sorg 1997, 2002; Knüsel 2010; Pokines and Symes 2013). Taphonomic approaches have been applied in studies ranging from the effects of earthworms on archaeological site stratigraphy (Canti 2003) to the excavation and analysis of victims of modern war crimes (Simmons 2002). Of particular relevance to mortuary studies and Andean funerary archaeology is the publication of many actualistic, archaeothanatological, and empirical studies examining the influence of natural patterns of decomposition and decay (including taphonomic agents such as necrophagous and osteophagous insects and carnivores) and the roles humans play in the postmortem fates

6 6   Backo

of human remains, especially the ritualized and physical ways the living interacted with their dead (Backo 2012; Duday 2009; Huchet and Greenberg 2010; Huchet et al. 2011; Klaus and Tam 2015; Millaire 2004; Rakita et al. 2005). Here, the conception of taphonomy involves the study and description of all postmortem processes, both natural and culturally derived, affecting the subsequent recovery, condition, and interpretation of skeletal remains. Following biological death, a human body typically undergoes a characteristic sequence of decomposition and disarticulation until complete skeletonization occurs. Further weathering and breakdown will eventually reduce bone until it is either completely resorbed back into the ecosystem or becomes fossilized. Elevated heat and humidity are primary environmental accelerators of decomposition, while necrophagous insects are the primary external accelerators (Bass 1997; Campobasso et al. 2001; Ortner et al. 1972; Rodriguez and Bass 1983). A comparison of two studies, one conducted in an arid desert environment akin to that found on the north coast of Peru (Galloway 1997), and the other conducted in the more humid, temperate climate of Tennessee (Bass 1997), illustrates the variability that can be found among decay and disarticulation rates. Each study found that decomposition occurred in five stages, culminating with skeletonization and eventual breakdown of exposed elements. Remains in the hotter and drier desert environment passed through each successive stage at a faster rate than bodies in Tennessee. Exposure of the bones was usually reached by two months in both settings, followed by sun bleaching and exfoliation of cortical bone. Galloway (1997), however, inserted an additional stage found in desert environments—mummification—which could delay skeletonization by several days to several months. Archaeologists and bioarchaeologists may examine mortuary assemblages to determine the interval between death and deposition, define demographic structures, and identify the physical remnants of ritual activities. Unusual depositional and environmental conditions, especially in the cases of commingled, exposed, or modified burials, generate complex observational, analytical, and interpretive challenges (Adams and Byrd 2008, Backo 2012, Duday 2009; Ubelaker 1974, 2008). In modified assemblages, analytical issues extend to include the identification of biotic, abiotic, and culturally mediated agents of alteration. Such forces can bias skeletal element representation and recovery rates, alter disarticulation patterns, produce pseudopathological bone damage or body positioning, and relate to deliberate versus random forces acting

The Taphonomy of Ritual Killing  67

on mortuary contexts. Complicating factors include activities involving the selective removal of elements, the disarticulation, transport, and subsequent deposition of remains, and ritualistic placement. The first two factors result from both faunal and human activities, as well as environmental agents such as fluvial action, whereas the final factor is a uniquely human activity. The ability to confidently distinguish between natural and culturally mediated influences is critical to the reconstruction of events for any site with a complicated depositional history. While scientific burial excavations have been conducted in Peru for more than a century, a long-standing focus on an art history paradigm and the sampling of gravelots left skeletal remains generally sidelined. Explicit and detailed taphonomic approaches in Andean funerary archaeology have remained long underdeveloped (Klaus 2011). Work by Nelson (1998) at San José de Moro and Verano (1995, 1997b) at Sipán were the first to capitalize on taphonomy and shed light on then-unrecognized practices of delayed primary burial in Moche funerary customs. Subsequent taphonomic research by Millaire (2004), Klaus (2008a, 2011), Cordy-Collins (2009), Backo (2012), and Klaus and Tam (2015) have further added to reconstructions of north coast funerary rituals. Still, natural and cultural forces acting in the formation of sacrificial contexts specifically have remained understudied.

Materials and Methods Archaeological and Depositional Contexts The contexts selected for this analysis were Huaca de la Luna’s Plaza 3A and the mass grave deposited outside the walls of the Huaca 1 Complex at Pacatnamú (Verano 1986a). These assemblages were chosen due to evidence of traumatic death and subsequent unusual treatment of the remains, which included evidence of exposure, mutilation, defleshing, and dismemberment (Verano 2001a, 2008a). Each assemblage also appeared to be affected by a wide variety of potential natural and cultural taphonomic agents. In addition, each depositional setting consisted of a distinct disposal space, separate and demarcated from other secular or ritual activities. Huaca de la Luna Plaza 3A Huaca de la Luna is a large adobe ceremonial complex located at the base of Cerro Blanco. It is one of two principal adobe brick mounds at the site of Huacas de Moche in the Moche river valley, which consists of three

6 8   Backo

large truncated pyramidal structures (Huaca del Sol, Huaca de la Luna, and the New Temple/Huaca de la Luna Platform III), and an urban zone with associated cemeteries. The site appears as the focal point of Moche culture during the Moche III–IV periods (ca. AD 300–550). It is also known for dramatic evidence of human sacrifice (Bourget 1997, 1998a, 1998b; Verano 2000; Verano et al. 2008; Hamilton, chapter 2; Verano and Phillips, chapter 9). Huaca de la Luna Plaza 3A (figure 3.1) dates to the late Moche IV era (Bourget 2001a). It covers 1,930 m2 and deliberately encloses a rocky outcrop at the base of Cerro Blanco. Excavation of Plaza 3A by Steve Bourget in 1995–1996 recovered a large quantity of human remains. All were males in the adolescent to middle adult age range (15–39 years) indicating a high degree of specificity in victim selection (Verano 2000, 2001a). Common perimortem injuries involved cut marks indicative of deep lacerations to the throat in addition to blunt force cranial trauma (Hamilton, chapter 2; Verano 2000). The remains appear to have been deposited in several distinct but relatively closely timed events. Each was separated stratigraphically by alternating layers of hard clay and windblown sand, indicating that the killings coincided with at least two periods of heavy rainfall that washed

3.1. The sacrificial and depositional setting of Huaca de la Luna Plaza 3A, today filled in by desert sands. In 1995, Steve Bourget encountered skeletal remains of Moche sacrifice victims here. Platform II is directly behind the rocky outcrop. Photo by Haagen Klaus.

The Taphonomy of Ritual Killing  6 9

3.2. Huaca de la Luna Plaza 3A sacrifice victims, Strata 1 (DS2, or Dentro Sedimento 2). Based on original field maps courtesy of Steve Bourget and redrawn by Haagen Klaus.

3.3. Huaca de la Luna Plaza 3A sacrifice victims, Strata 2 (SS2, or Sobre Sedimento 2). Based on original field maps courtesy of Steve Bourget and redrawn by Haagen Klaus.

7 0  Backo

clay down into the plaza from the adobe walls (Bourget 1997, 1998b). The dead were haphazardly deposited within the soft clay and sand and appear to have been then exposed on the plaza floor. Bleaching and weathering of the bones, as well as the presence of numerous fly puparia, indicate prolonged exposure to the elements (Bourget 1998b; Verano 1998). Plaza 3A was a highly complex context, strewn with more than two thousand isolated bones and articulated body parts along with more than a dozen recognizable, though incomplete, individual skeletons. Bourget (1997, 1998b) documented most of the human remains as isolated, disarticulated bones, but the deposit also included numerous articulated and semi-articulated (though disembodied) elements (figures 3.2, 3.3); single bones are present or in close association with non-contiguous anatomical elements, such as a bony sacrum adjacent to a humerus, or a cranium atop a tibia. Semi-articulated elements were found in other cases, such as an isolated but complete pelvic girdle, multiple relatively complete limbs, and sections of articulated vertebral columns. Grouping of elements was also observed, such as with clusters of disarticulated crania and articulated skulls found together, isolated mandibles facing each other, and stacked and paired disarticulated long bones. Complete skeletons, however, were rare. The most coherent sets of remains corresponded at most to 15 individuals. Most of the remains can be classified as isolated elements, groups of articulated elements, or, less commonly, skeletons missing elements. Evidence for deliberate modification or disarticulation in the form of sharp force trauma, however, is not present in this sample to any significant degree (Hamilton, chapter 2; Verano 1998, 2000). Taphonomic explanations must be explored to understand the formation process(es) at play in Plaza 3A. The Pacatnamú Mass Burial Close to the end of the Moche IV period (AD 500–600), a climatic and social upheaval caused a political shift northward, away from Huacas de Moche to Pampa Grande in the Lambayeque valley (Shimada 1994a; Shimada et al. 1991). Ritual killing continued into the Late Intermediate Period at Pacatnamú in the Jequetepeque river valley. Spanning Moche, Lambayeque/Sicán, and Chimú occupations from AD 550–1350, Pacatnamú was located at the mouth of the Jequetepeque River and was a paramount administrative-ceremonial center in the lower valley. The principal ceremonial complex of Huaca I was constructed at

The Taphonomy of Ritual Killing  7 1

3.4. Composite view of the Pacatnamú mass grave just outside of the Huaca 1 complex. Based on original drawings by Genaro Barr from Verano (1986) and redrawn by Haagen Klaus.

7 2   Backo

Pacatnamú between AD 1100–1400, overlapping almost entirely the Late Lambayeque (or Late Sicán) provincial era (Donnan 1986a). Multiple killings took place here around the time of the Chimú conquest of the Jequetepeque valley (Verano and Phillips, chapter 9). The remains of at least 14 adolescent to middle adult male individuals were unceremoniously deposited at the bottom of a nearly 3 meter–deep defensive trench adjacent to the entrance to Huaca 1 (figure 3.4) (Verano 1986a). Stratigraphic evidence indicated these bodies accumulated over at least three events, separated by layers of sand and rubble. Weathered bone surfaces and extensive deposits of fly puparia (Faulkner 1986) indicated the victims were left exposed to the elements, and remains of rope fragments used to bind or hobble the victims were also found. Extensive perimortem trauma in the form of fractures and cut marks was consistent with stab wounds, decapitation, and catastrophic blunt force cranial trauma. Five individuals had an intriguing pattern of cut marks and fractures on the ribs and sternum attributed to opening of the chest cavity (Verano 1986a). This pattern links to recent evidence from the New Temple at Huacas de Moche, indicating that chest-opening developed there during Moche V, tying the two sites and cultures together (Backo and Verano 2013; Verano and Backo, in press). In addition, members of each stratigraphically separated group at Pacatnamú revealed a pattern of missing elements. In the first group, all four individuals were missing the left radius. Cut marks and perimortem fractures in the wrist and elbow area indicate these bones were deliberately removed around the time of death. In the second and third groups, some individuals were missing a left leg or arm (Verano 1986a). Taphonomic Methods The depositional contexts at Huaca de la Luna Plaza 3A and the mass burial at Pacatnamú represent complex deposits of human remains associated with ritual violence. Detailed taphonomic reconstructions of these settings can provide more complete and holistic understandings of ritual violence on the north coast—especially by contributing perspectives on sacrifice as a ritual process, spanning pre-, peri-, and post-sacrificial acts. Here, accurate estimations of original sample size, element recovery and representation, and anatomical mapping of disarticulation patterns help open new windows on postmortem treatment in settings of ritual killing.

The Taphonomy of Ritual Killing  7 3

Accurate Estimation of Original Sample Size Estimating how many individuals are actually represented in the assemblage is invaluable for demographic profiles and potential identifications in forensic investigations or archaeological settings, laying the groundwork to calculate additional parameters (see below). An accurate estimation of sample size may not be straightforward if a majority of individuals are scattered into separate elements or are commingled. The most common estimator for determining the number of individuals within a sample is the minimum number of individuals (MNI) technique, where a count of each identified and sided element is translated into the minimum number of individuals that could then correspond to the identified elements (Orchard 2005; White 1953). This method is simple, easily replicated, and lacks overestimation bias (Adams and Konigsberg 2008; Ubelaker 2002). Despite its popularity, MNI is not necessarily the most accurate estimator of original sample size, especially in highly commingled and modified collections. It represents, by definition, only the minimum number of individuals (Adams and Konigsberg 2004, 2008; Fieller and Turner 1982; Grayson 1978; Marean et al. 2001; Orchard 2005). MNI values can be negatively biased in assemblages where fragmentation is high or where bones are scattered, disturbed, broken, or commingled. Such conditions may result from various kinds of funerary activities, natural processes, or a combination of both (Mundorff et al. 2008; Ubelaker 1974). The Lincoln Index (LI) and Chapman Estimator (CE) attempt to correct for such bias (Adams and Konigsberg 2004). However, these equations require pair matches between corresponding elements (i.e., left and right radius from the same individual). These techniques also require that data loss be randomly distributed, which is only rarely found in scavenging or ritual behaviors. The LI also tends to underestimate original assemblage size when samples are small (Adams and Konigsberg 2004). Another alternative considered by Adams and Konigsberg (2004) is the most likely number of individuals (MLNI) equation. Derived from the Lincoln Index, MLNI accounts for the recovery probability of an element or matched pair; biases are reduced, and a value is calculated that supplements MNI and LI techniques. This method is especially suited for commingled or unassociated elements (i.e., mass burials, ossuaries). Despite the demographic homogeneity of these samples, there was

74   Backo

enough variation in stature, robusticity, and age to confidently establish pair matches for humeri and femora at Plaza 3A and Pacatnamú. Each pair match was peer-reviewed in the field by a second physical anthropologist. Additional confirmation of some matches was fortuitously achieved when bones incorrectly curated as an isolated element were found to match to an individual with whom it was originally excavated. Each technique possesses various strengths and weaknesses. So, during the analysis of the Pacatnamú remains, a decision was made to create a new index—a straightforward average of all the estimators. This approach helps to balance biases between various estimators and counter the possibility of over- or underestimating the original assemblage size. This final estimator was designated the derived number of individuals (DNI). The DNI is a conservative sample size estimator created by calculating, and then averaging, the MNI, MLNI, and LI. This value is then compared with the number of paired bones with sided, but isolated, unpaired bones (which may each represent one unique individual). As an example, the total number of femora collected from Pacatnamú was 41. Thirty-four of these bones were paired together, creating 17 pairs of left and right elements; the seven remaining bones could each represent a single additional individual. When this value is added to the number of pairs (each pair representing one unique individual), the number obtained is 24 (17 + 7 = 24). This estimate is often referred to as the Grand Minimum Total (GMT) (Adams and Konigsberg 2008), and as noted later in this chapter, it is identical to values obtained by the DNI for Pacatnamú. The same result was also obtained when comparing paired and non-paired humeri, further illustrating the strength of the DNI as a robust estimator. The element best suited for computing the DNI appears to be the paired element closest in number to the MNI. Element Recovery and Representation With sample size estimates established, the degree of completeness of skeletal remains can be calculated. This approach compares the number of bones or elements that should be present given the estimated sample size versus how many of such elements were, in fact, represented in the original assemblage. This difference can be calculated as a percentage, with loss values tested for patterns of statistical significance using the χ2 test for individual elements. Bone size and mass are key factors that influence rates of recovery and representation. With all factors being theoretically equal, larger bones (such as crania and femora)

The Taphonomy of Ritual Killing  7 5

are recovered in greater numbers than smaller elements (Lyman 1994) such as hand and toe phalanges, carpal and tarsal bones, or sesamoid bones—which have a greater propensity to separate from the body rapidly (Duday 2009). Further, the use of the recovery probability function (r) calculates the likelihood that a certain bone type will be recovered, assuming that both right and left sides of an element have equal probability of representation (r = 2P/L+R) (Adams and Konigsberg 2004). For the recovery of any element to be considered high, r-values should be 0.70 or greater. Anatomical Mapping of Disarticulation Patterns To characterize patterns of disarticulation in comparative perspective, it is necessary to translate the presence or absence of an articulated joint into a value that can be compared between elements and sides. Each element with a preserved articular surface was scored as articulated or disarticulated. These were then correlated to field maps, notes, and curated associated elements to determine if that joint was recovered in an articulated or disarticulated state. These data were then used to calculate joint ratio scores. Joint ratios are calculated as the number of articulated joints to the number of disarticulated joints, with scores obtained by dividing the former by the latter. Ratio scores greater than 1.0 include all joints for which there are a greater number of articulated than disarticulated specimens. Ratio scores between 0.99 and 0.0 include those joints that were more disarticulated. The mean, the mean minus the two most outliers, median, standard deviation, and Z-scores can be computed from ratio scores if desired. Joint ratio scores are ordered from lowest and most disarticulated, to highest and most intact. The goal of this process is to compare entire patterns of missing bones across contexts through which natural and cultural taphonomic agents can be evaluated (i.e., butchery related to food consumption, carnivore-scavenged remains). This is further facilitated by translating quantitative data into a qualitative image to discern similarities and differences within or between temporal or spatial contexts. Translating joint scores into colors (or by a similarly effective visual convention) onto a standardized skeletal image presents patterns visually to enable direct comparisons. Here, four relative shaded categories were created: medium dark gray is the most disarticulated, followed by light gray, medium gray, and dark gray, the latter representing the most complete joints.

7 6   Backo

Results Accurate Estimation of Original Sample Size The original estimated number of individuals represented in Huaca de la Luna Plaza 3A was 70+ individuals; this was based on a count of larger, resilient long bones. The original analyses used an approximate value rather than an absolute count (Verano 2000). Here, MNI was examined again but included a count for every recovered element (table 3.1). The results show a range of values, with most clustering around an MNI value between 50–60, but they spanned from as low as 30 (distal Table 3.1. MNI estimations from Huaca de la Luna Plaza 3A and the Pacatnamú Mass Grave Skeletal Assemblage Element Type/ Element Group

N Observed Elements, Huaca de la Luna Plaza 3A

N Observed Elements, Pacatnamú Mass Grave

Cranium

47

17

Hyoid Bone

1

7

Mandible

62

11

Cervical Vertebrae

64

13

Thoracic Vertebrae

67

14

Lumbar Vertebrae

72

15

Right Ribs

64

14

Left Ribs

62

14

Body of the Sternum

65

11

Left Clavicle

63

14

Right Clavicle

74

13

Left Scapula

60

15

Right Scapula

66

15

Left Humerus

72

15

Right Humerus

73

15

Left Ulna

76

17

Right Ulna

61

14

Left Radius

71

10

Right Radius

65

10

Carpals

48

12

The Taphonomy of Ritual Killing  7 7

foot phalanges) to as high as 76 (for the proximal left ulna). Thus, 76 represents the new MNI estimate for Plaza 3A. The new MNI calculated for the Pacatnamú mass grave is 21 (table 3.1), based on the counts of proximal right femora. There is less variability of MNI estimates between the skeletal elements at Pacatnamú, which is undoubtedly a function of far lower degrees of commingling and scattering of remains. Next, alternative estimators (LI, MLNI, CE, and DNI) were examined first by pair-matching humeri and femora. Humeri and femora were chosen due to their nature as large, conservable elements with a high recoverability rate. A total of 13 pairs, out of a possible total of 15, was Table 3.1. Continued.

Element Type/ Element Group

N Observed Elements, Huaca de la Luna Plaza 3A

N Observed Elements, Pacatnamú Mass Grave

Metacarpals

55

14

Proximal Hand Phalanges

51

13

Distal Hand Phalanges

37

9

Sacrum

73

15

Left Os Coxae

69

16

Right Os Coxae

67

15 20

Left Femur

60

Right Femur

59

21

Patellae

60

12

Left Tibia

68

17

Right Tibia

64

18

Left Fibula

66

16

Right Fibula

67

17

Left and Right Talus

76

15

Left and Right Calcaneus

74

16

Left and Right Tarsals

57

14

Left and Right Metatarsals

63

16

Proximal Foot Phalanges

51

9

Distal Foot Phalanges

30

5

MNI Estimate

76

21

7 8   Backo Table 3.2. Alternative Estimators of Total Sample Size for Huaca de la Luna and Pacatnamú, Including the Lincoln Index (LI), Chapmans Estimator (CE), and the Most Likely Number of Individuals (MLNI) + DNI (Derived Number of Individuals) Huaca de la Luna LI: NLI = (LxR)/P Plaza 3A

CE: [(L+1)(R+1)/P+1-1]

MLNI: [(L+1)(R+1)-1/P+1–1] MNI/DNI

Humerus

117

117

117

76/107

Femur

89

89

89

76

Pacatnamú

LI: NLI = (LxR)/P

CE: [(L+1)(R+1)/P+1-1]

MLNI: [(L+1)(R+1)-1/P+1–1] MNI/DNI

Humerus

17

18

18

21

Femur

25

25

25

21/24

Table 3.3. Recovery probabilities for Huaca de la Luna Plaza 3A and the Pacatnamú Mass Grave Site

Element

Recovery Probability (r = 2P/L+R)

Huaca de la Luna Plaza 3A

Humerus

0.62

Femur

0.67

Pacatnamú Mass Grave

Humerus

0.87

Femur

0.829

made for the humeri, and 17, out of a possible total of 20, for the femora of Pacatnamú. For Plaza 3A, however, 137 humeri were complete enough for pair-matching, and only 45 matches were made. For the femora, a total of 118 bones were complete enough, with 40 pair matches resulting. Utilizing these data (table 3.2), the LI calculated for Plaza 3A is 117 based on the humerus and 89 based on the femur. Identical values are obtained for both the CE and the MLNI. The LI at Pacatnamú was calculated as 17 for the humerus and 25 based on the femur. CE and MLNI values were 18 for the humerus and again 25 for the femur. This higher degree of consistency between the MNI and alternative estimators again underscores the greater degree of skeletal completeness at Pacatnamú compared to Plaza 3A. While computation of the DNI value for Pacatnamú is straightforward, DNI may be skewed in the Plaza 3A sample given the low recovery

The Taphonomy of Ritual Killing  7 9

probability of the humerus and femur (table 3.3). Ulnae, which provided the final MNI count, were not pair matched due to time constraints. Yet absolute counts of the humeri are close in number to the ulnae, and the humeri were used as a proxy to compute DNI and determine the final sample size estimate at Plaza 3A. Using the humerus, the range obtained by the four methods was between 76 and 117. If the average of these values is taken, the DNI for Plaza 3A is 107 individuals. For Pacatnamú, the DNI value derived from the femur is 24 individuals. Element Recovery and Representation Based on a sample size of 107 individuals, percentage loss (PL) was calculated for the skull, long bones, sternum, and phalanges at Huaca de la Luna Plaza 3A (table 3.4). This loss ranges from a low of 29% for the left ulna, to a high of 72% for the distal foot phalanges. The right and left humerus and the left ulna were elements characterized by a low PL. Elements with a high rate of loss compared to the average (greater than 10%) include the distal hand and foot phalanges and the cranium. Chisquare tests indicate that the loss pattern of the cranium is statistically significant at the 0.05 level, and foot and hand phalanges loss is significant at the 0.01 level. This high rate of loss for phalanges can be explained by their small size. Size influences rates of recovery for elements, with larger elements being recovered at higher numbers than smaller elements in collections subjected to natural taphonomic agents (Lyman 1994). If this rule of size is taken into account, however, there are still elements with unusually high or low rates of loss compared to that expected. In particular, the cranium and the femur stand out as having a higher rate of loss than expected (table 3.6). Both these elements are characterized as large bones and so should be expected to have low rates of loss. Table 3.5 demonstrates the loss values for skeletal elements by size at Pacatnamú. The average percent loss data for Pacatnamú reveal a different pattern, whether considered by element (table 3.5) or bone size (table 3.6). Loss values range from a low of 12% for the right femur to 79% for the distal foot phalanges. However, most values cluster in the 30% loss range, with the significant exception of the right and left radius (PL = 58%; p = 0.01). Recovery probability function values (r) also differ between Pacatnamú (rhumerus =0.87; rfemur = 0.83) and Plaza 3A (rhumerus = 0.62; rfemur = 0.67) (Backo 2012).

8 0  Backo Table 3.4. Element Rates of Loss, Huaca de la Luna Plaza 3A Element

MNI

Difference, MNI-DNI (107)

Percentage loss

Cranium

47

60

56%*

Mandible

62

45

42%

Left Humerus

72

35

33%

Right Humerus

73

34

32%

Left Radius

71

36

34%

Right Radius

65

42

39%

Left Ulna

76

31

29%

Right Ulna

61

46

43%

Left Clavicle

74

33

31%

Right Clavicle

63

44

41%

Proximal Hand Phalanges

51

56

52%

Distal Hand Phalanges

37

70

65%**

Body of the Sternum

65

42

39%

Left Femur

60

47

44%

Right Femur

59

48

45%

Left Tibia

68

39

36%

Right Tibia

64

43

40%

Left Fibula

66

41

38%

Right Fibula

67

40

37%

Proximal Foot Phalanges

51

56

52%

Distal Food Phalanges

30

77

72%**

* Significant at the 0.05 level.

** Significant at the 0.01 level.

Anatomical Mapping of Disarticulation Patterns Joint ratio scores are shown in tables 3.7 and 3.8. These can then be separated into sequential groups, obtained by artificially dividing the chart into four more or less equal groups. Group 1 represents joints with the least number of intact articulations, while Groups 2, 3, and 4 are the joints that are progressively and relatively less disarticulated and complete. At Plaza 3A, elements with the highest joint ratio scores (over one standard deviation from the mean) and the greatest number of preserved articulations include the elements of the pelvic girdle, lumbar,

The Taphonomy of Ritual Killing  8 1 Table 3.5. Element Rates of Loss, Pacatnamú Mass Grave Element

MNI

Difference, MNI-DNI (24)

Percentage Loss

Cranium

17

7

29%

Mandible

11

13

54%*

Left Humerus

15

9

37%

Right Humerus

15

9

37%

Left Radius

10

14

58%**

Right Radius

10

14

58%**

Left Ulna

17

7

29%*

Right Ulna

14

10

42%

Proximal Hand Phalanges

13

9

37%

Distal Hand Phalanges

9

15

62%**

Left Femur

21

4

17%**

Right Femur

21

3

12%**

Left Tibia

17

7

29%*

Right Tibia

18

6

25%*

Left Fibula

16

8

33%

Right Fibula

17

7

29%*

Proximal Foot Phalanges

9

15

62%**

Distal Food Phalanges

5

19

79%**

* Significant at the 0.05 level

** Significant at the 0.01 level

Table 3.6. Rates of Bone Loss Categorized by Bone Size, Huaca de la Luna Plazas 3A and 3C, and the Pacatnamú Mass Grave A. Bone Size Category

Plaza 3A

Plaza 3C

Pacatnamú Mass Grave

Average (excluding outliers)

Small

50%

47%

55%

51%

Medium

37%

35%

42%

36% (excluding Pacatnamú)

Large

44%

24%

22%

23% (excluding 3A)

B. Element Category

Observed Percent Loss

Expected Percent Loss

X2

Plaza 3A Cranium

56%

22%

52.55**

Plaza 3A Left Femur

44%

22%

22.0**

Plaza 3A Right Femur

45%

22%

24.05**

* Significant at the 0.05 level

** Significant at the 0.01 level

2   la Backo Table 3.7. Joint Ratio Scores, Huaca8de Luna Plaza 3A Articulation Cranium-Mandible

Articulated Disarticulated 17

90

Ratio 17:90

Ratio score Group 0.19

1

Cranium-C1 Vert.

14

93

14:93

0.15

1

C1-C2 Vert.

24

83

24:83

0.29

1

Cervical Verts.-Cervical Verts.

46

61

46:61

0.75

3

Thoracic Verts.-Thoracic Verts.

54

53

54:53

1.02

4

Lumbar Verts.-Lumbar Verts.

55

52

55:52

1.06

4

L5 Vert.-Sacrum

46

61

46:61

0.75

3

L. Scapula-Clavicle

34

73

34:73

0.47

2

Rt. Scapula-Clavicle

35

72

35:72

0.48

2

L. Clavicle-Sternum

35

72

35:72

0.48

2

Rt. Clavicle-Sternum

35

72

35:72

0.48

2

L. Humerus-Scapula

30

77

30:77

0.39

2

Rt. Humerus-Scapula

30

77

30:77

0.39

2

L. Humerus-Radius

30

77

30:77

0.39

2

Rt. Humerus-Radius

23

84

23:84

0.27

1

L. Humerus-Ulna

33

74

33:74

0.45

2

Rt. Humerus-Ulna

26

81

26:81

0.32

1

L. Ulna-Radius

45

62

45:62

0.73

3

Rt. Ulna-Radius

38

69

38:69

0.55

3

L. Radius-Carpals

37

70

37:70

0.53

3

Rt. Radius-Carpals

29

78

29:78

0.37

1

L. Ulna-Carpals

38

69

38:69

0.55

3

Rt. Ulna-Carpals

24

83

24:83

0.29

1

L. Os Coxae-Sacrum

47

60

47:60

0.78

3

Rt. Os Coxae-Sacrum

44

63

44:63

0.70

3

L. & Rt. Os Coxae

51

56

51:56

0.91

4

L. Femur-Os Coxae

33

74

33:74

0.45

2

Rt. Femur-Os Coxae

33

74

33:74

0.45

2

L. Femur-Tibia

29

78

29:78

0.37

1

Rt. Femur-Tibia

35

72

35:72

0.48

3

L. Tibia-Fibula

47

60

47:60

0.78

4

Rt. Tibia-Fibula

51

56

51:56

0.91

4

L. & Rt. Fibula-Talus

66

41

66:41

1.61

4

L. & Rt. Fibula-Talus

70

37

70:37

1.89

4

L. & Rt. Talus-Calcaneus

77

30

77:30

2.57

4

The Taphonomy of Ritual Killing  8 3

and thoracic vertebral columns. This is not unexpected, as these segments are characterized by large amounts of tissue and possess some of the strongest and most persistent joints of the skeleton (Duday 2009; Haglund et al. 1989; Toots 1965). For Plaza 3A, however, the element with the lowest score, and therefore least number of intact articulations, is the skull. This includes articulation with the mandible, and with the C1 cervical vertebra. At Pacatnamú, the most highly disarticulated structures are the radius and knee (table 3.8). Table 3.8. Joint ratio scores, Pacatnamú Articulation

Articulated

Disarticulated

Ratio

Ratio score

Group

Cranium-Mandible

10

6

5:3

1.67

2

Cranium-C1 Vert.

10

6

5:3

1.67

2

C1-C2 Verts.

11

1

11:1

11

4

Cervical Verts.-Cervical Verts.

11

2

11:2

5.5

3

Thoracic Verts.-Thoracic Verts.

11

4

11:4

2.75

3

Lumbar Verts.-Lumbar Verts.

14

1

14:1

14

4

L5 Vert.-Sacrum

15

0

15:0

15

4

Clavicle-Scapula1

22

8

11:4

2.75

3

Sternum-Rt. Clavicle

11

2

11:2

5.5

3

Sternum-L. Clavicle

9

4

9:4

2.25

3

Humerus-Rt. Scapula

13

2

13:2

6.5

3

Humerus-L. Scapula

10

5

2:1

2

3

Humerus-Radius1

16

13

16:13

1.23

1

Humerus-Ulna1

22

9

22:9

2.44

2

Ulna-Radius1

17

14

5:2

1.21

1

Sacrum-Os Coxae1

28

3

28:3

9.33

4

Os Coxae-Os Coxae

14

2

7:1

7

3

Os Coxae-Femur1

26

15

26:15

1.73

2

Femur-Rt. Tibia

13

8

13:8

1.62

1

Femur-L. Tibia

11

7

11:7

1.57

1

Tibia-Fibula

21

14

3:2

1.5

1

Tibia-Talus

22

12

11:6

1.83

2

Fibula-Talus

19

13

19:13

1.46

1

Talus-Calcaneus

26

3

26:3

8.67

4

1

Combined counts for both the left and right side.

8 4   Backo

Discussion The results from these two contexts indicate the fates of the ritualistically killed were not only very complex but also highly variable. Patterns of disarticulated skeletons provide clues to these processes and the meanings and ideas regarding the treatment of victims after death. Disarticulation and Taphonomic Winnowing Huaca de la Luna Plaza 3A In both settings, sample size estimates and percent loss data show the varying ways in which bodies were either disarticulated or maintained in a relatively intact state. Some missing bones and disarticulated joints may be expected via natural processes such as scattering and loss of distal hand and foot phalanges. Crania and femora, however, were clearly underrepresented, indicating individuals deposited in Plaza 3A were subjected to a higher degree of taphonomic winnowing than those at Pacatnamú. What factors, then, were responsible for this pattern in Plaza 3A? Carnivore scavenging is a well-documented natural taphonomic force that can influence element representation in exposed remains. Carnivores are not random in their feeding behaviors but produce telltale patterns of dismemberment, element destruction, and trauma. To better define these patterns, disarticulation, element representation, and damage patterns were extracted from a wide variety of forensic anthropological and archaeological reports (Carson et al. 2000; Haglund 1992, 1997; Haglund et al. 1988, 1989; Haynes 1980, 1983; Hill 1979; Kuckelman et al. 2002; Marshall 1989; Milner and Smith 1989; Murad and Boddy 1987; Toots 1965). A synthesis of damage and recovery rates is presented in figure 3.5. Carnivores remove elements as they are consumed or transported to another location and tend to process a body in a predictable sequence resulting in certain elements (such as the upper limb and shoulder girdle) recovered in association (Haglund et al. 1988, 1989; Hill 1979). However, this can be ruled out in the case of Plaza 3A, as both the femur and the cranium are traditionally well represented in carnivore-scavenged remains. Such large bones are not easily destroyed or transported away by small to midsized carnivores on the north coast of Peru (domestic dogs, vultures). For carnivore-scavenged

The Taphonomy of Ritual Killing  8 5

3.5. Graph illustrating comparative damage and recovery rates of individual elements obtained through published data of faunal and forensic remains subjected to exposure and carnivore activity. Image by Heather Backo.

remains, the cranium has the highest reported recovery rate of all. Also, not a single cranium or femur was recovered from Plaza 3A with damage attributable to carnivore activity. Insect activity and sunlight-related bleaching indicates these remains were exposed for prolonged periods, but they clearly appeared to have been protected from scavengers. Plaza 3A was subjected to periodic heavy rainfall around AD 550. Fluvial activity may have contributed to bone transport and sorting. Water transport affects elements primarily according to their size and density and secondarily by shape. Bones subjected to hydraulic sorting illustrate groupings of elements by these characteristics, rather than the anatomical groupings found within an articulated skeleton (Lyman 1994). The rounded cranium generally falls into the high transport category, while the femur consistently falls into the low transport (or drag) category. Fluvial activity would also not tend to act on the femur alone without also transporting the tibia (Behrensmeyer 1975; Boaz and Behrensmeyer 1976; Lyman 1994; Nawrocki et al. 1997; Voorhies 1969). Fluvial transport therefore can also be ruled out as the primary force shaping the Plaza 3A assemblage. The taphonomic agent(s) affecting element representation in Plaza 3A is not likely natural. Humans commonly use knives to disarticulate

8 6   Backo

animals or other humans, but there are surprisingly few cut marks visible on the elements from Plaza 3A that can be attributed to disarticulation (Verano 2000). Of all the bones, only 19 show sharp force trauma definitely related to dismemberment, representing 12% of all elements with cut marks. This contrasts to the evidence at Pacatnamú, where 92% of cut marks were associated with dismemberment (Backo 2012). The low recovery levels of the cranium could be attributable, at least in part, to perimortem trauma that caused extensive destruction of the cranial vault for many individuals. Yet there was also a high degree of perimortem damage to crania at Pacatnamú without a low recovery rate. There was minimal perimortem trauma associated with the femur disarticulation in Plaza 3A. Only one femur possessed cut marks attributable to disarticulation, and one showed perimortem trauma in the form of a spiral fracture. Therefore, it appears as though perimortem trauma alone, in particular sharp force, should not be relied upon to indicate the degree of element collection for Plaza 3A. The Pacatnamú Mass Burial While this setting of victim disposal was also characterized by complex taphonomy, the patterns of bone representation and winnowing at Pacatnamú is distinct from Huaca de la Luna. Calculations of PL based on a sample size of 24 individuals ranges from a low of 12% (right femur) to a high of 79% (distal foot phalanges). Statistically significant losses in this sample include the mandible, the right and left radius, and the left ulna, while lower rates of loss characterized the right and left femora, the right and left tibia, the right fibula, all foot phalanges, and distal hand phalanges. Representation patterns of several elements in the Pacatnamú assemblage cannot be explained simply by size and durability. The mandible is characterized by a fairly high rate of loss (54%) when compared to the cranium (29%). The right and left radius demonstrate a very high loss rate for their size category (58%). This rate, in fact, is greater than that of the hand phalanges. This suggests an artificial agent has also affected element representation at Pacatnamú. Carnivore activity, again, could promote loss of certain elements as they are removed for consumption at another location. However, carnivoreravaged assemblages lack this degree of single-element specificity and intentional targeting of bones for removal (figure 3.7). As noted in figure 3.5, the humerus, radius, and ulna all share a similar damage and recovery rate, indicating that carnivore activity tends to act similarly on the upper limb as a whole. Carnivore-related cases also demonstrate

The Taphonomy of Ritual Killing  87

comparable recovery and damage rates for the cranium and mandible, the latter of which has high recovery. Fluvial activity would not tend to act on the radius separately from the ulna, as they share a similar shape, density, size, and common muscular and ligamentous attachments. The mandible is a low transport element (Boaz and Behrensmeyer 1976; Nawrocki et al. 1997; Voorhies 1969), and its high percentage loss here is inconsistent with fluvial mechanisms. This too leaves cultural activities as the most likely cause for the high rate of loss of both radii and mandibles at Pacatnamú. Cut mark data from Pacatnamú is more informative than at Plaza 3A, however. Ninety-two percent of the documented cut marks were related to dismemberment. Of particular significance are cuts on ulnae and carpal bones. Out of 31 recovered ulnae, five had cut marks on their distal aspects adjacent to the distal radioulnar joint and the carpal bones of the wrist. One ulna had multiple circumferential cut marks encircling the distal end. Three sets of carpal bones showed cut marks, all of which were located on bones associated with the articulation of the distal radius with the wrist. A

b

c

3.6. Examples of perimortem trauma to crania from Pacatnamú. (A) Anterior view of Cranium IB6 and inset (below) showing the cranial base with fractured mastoid process and fractured base of the zygomatic arch at center. (B) Anterior view of Cranium IB21 and inset (below) showing fractured mastoid process, fractured edges of missing cranial base, and a fractured zygomatic bone. (C) Anterior view of Cranium I and inset (below) detailing fractures to the left side of the cranium. Photos by Heather Backo.

8 8   Backo

Cut marks were absent, though, on mandibles and crania. All mandibles associated with articulated skeletons were in anatomical position (Verano 1986a). Only one individual lacked a mandible, so the remainder of disarticulation and loss would have occurred among the isolated crania present in the deposit. Each isolated cranium suffered extensive blunt force trauma, leading to the almost complete destruction of the vault or face (figure 3.6). Two plausible explanations for mandible loss at Pacatnamú exist: either mandible removal was facilitated by catastrophic trauma to the skull, or they were removed later as the decompositional process progressed. In either case, mandible removal was accomplished without the aid of sharp cutting tools. Ritual and Symbolic Factors Underlying Sacrifice Victim Taphonomy It appears that the living intimately interacted with the dead and decomposing bodies of victims of ritual violence over prolonged periods of time. In the Huaca de la Luna Plaza 3A sample, there were few to no cut marks attributable to disarticulation. This indicates that the living (perhaps ritual specialists) waited to a point when decomposition was sufficiently advanced to manually disarticulate the sacrifice victims. This raises novel implications for phenomenological and sensory aspects of living-dead interactions (sensu Meskell 1999) in contexts of sacrifice. The presence of intact articulations within disembodied elements (tibia-fibula, elbow joints, knee joints) indicates the decompositional process was not complete (Duday 2009) but advanced enough so that forceful pulling could separate a leg out of a hip joint, for example. At Pacatnamú, similar kinds of manual bone removal can be inferred, but there was also very clear use of cutting implements to remove elements from the upper limb specifically. Some skeletal elements could have been disturbed by individuals walking in Plaza 3A or as they deliberately separated or collected body parts. In such a scenario, joints would show high rates of disarticulation, including the easily levered hip and shoulder, followed by the elbow and knee; this matches the patterns illustrated in Plaza 3A. Also, the anatomical mapping process suggests that many of the limbs that remained in articulation were positioned at extreme angles of flexion, extension, or rotation anatomically impossible in life (figure 3.8). Bourget (1997, 1998b) theorized that these extreme positions represented individuals held down in the mud as they were executed and limbs were violently

The Taphonomy of Ritual Killing  8 9

carnivore pattern

huaca de la luna plaza 3a

pacatnamú mass grave

3.7. Disarticulation patterns for Huaca de la Luna Plaza 3A and Pacatnamú, compared with expected carnivore damage patterns. Group 1 (most disarticulated) = medium dark gray, Group 2 = light gray, Group 3 = medium gray, and Group 4 (least disarticulated) = dark gray. Image by Heather Backo.

twisted and pulled away from the torso. The findings here indicate it is equally possible that such positioning of these limbs occurred as a byproduct of very forcible postmortem dislocation and disarticulation. The taphonomic observations considered here help sort through the chaotic in situ nature of the remains in Plaza 3A. It points to a fairly consistent pattern of manipulation of heads and long bones by the living. It is quite difficult to say if the redeposition process held any tangible symbolic meanings or spatial significance. But in a few cases, bone placement was non-random, such as with the radial arrangement of three skulls at the foot of Individual E in Layer SS2, or with the nearby positioning of a pair of mandibles in dualistic opposition to each other (Bourget 1998b: fig. 54). Sample size estimators and percentage loss values converge on the fact that after disarticulation many bones simply went missing and evidently exited Plaza 3A forever. Heads and long bones (especially femora) seem to have been selectively targeted for such removal. More than half the original numbers of skulls are missing, just as more than a quarter of the original number of femora. Why this was done remains difficult to assess. Given the sacred and ritual character of these killings,

9 0  Backo

3.8. Skeletal remains of various individuals in Plaza 3A. The dislocation of the hip in both individuals may reflect postmortem manual manipulation. Photo by and courtesy of Steve Bourget.

the head may have been a “cathected object” that was “imbued with a powerful and efficacious spiritual essence” (Hill 2006:96; Weismantel 2015). Similar thinking may have surrounded the femur, a bone that in other Andean traditions held metaphors of fertility and life-generating properties (Salomon 1995; Klaus and Tam 2009a). Beyond the walls of Plaza 3A, the fate of some victims’ heads may be reflected in a pair of modified and curated adult male skulls documented in a niche located in the elite residential area adjacent to Huaca de la Luna (Verano et al. 1999). Similarly, cranial vaults recovered in Plaza 3A demonstrate perimortem chipping and semi-lunate chopping damage. These crania were likely being prepared either for display or use of some kind requiring intentional removal of the vault (figure 3.9). Similar cases of modified skulls were observed at Huaca de la Luna

The Taphonomy of Ritual Killing  91

Plaza 3C and at the New Temple (Verano and Backo, in press), while two carved cranial vaults were recovered from the Conjunto Arquitectónico 42 in the urban zone at Huacas de Moche (Seoane et al. 2008). At Huaca Cao Viejo in the El Brujo Complex (Chicama valley) an isolated femur was found incorporated into the base of a mural (Verano et al. n.d.). The femoral neck showed several circumferential cut marks, indicating it had been removed before connective tissue had decayed. Perhaps sacrifice victims’ bones were destined to become grave goods (sensu Hecker and Hecker 1992; also see examples [such as Burials M-IV-13, M-IV-25 at Moche] in Donnan and Mackey 1978). Others may have been destined for collection and then secondary burial at Huaca de la Luna (Uceda 1997). Missing bones from Plaza 3A may therefore indicate some victim body parts appear to have “lived on,” fulfilling other purposes in the ritual world of the Moche. Human sacrifice was a complex, multistage process; killing the victim was but one component. Bourget (2001b) argued that mandible removal might have served two purposes—to speed up decomposition, and to symbolically aid in the release of the victims’ anima, or spirit. The removal of other body parts, including cases of complete or nearly complete disincorporation, destruction, and scattering of bodies and body parts, likely goes beyond concepts of soul-transfer to the Moche afterworld. Careful, limited analogies between the Postclassic Maya (Duncan 2005) and the Moche suggest that the Moche process of tearing apart the rotting bodies of sacrifice victims may be rooted in expressions of destructive ritual violence. However, sacrifices in Plaza 3A may have necessitated such destruction. Body parts could have been an additional offering medium. Or, perhaps, disincorporation of the body or body parts was necessary to complete or terminate a sequence of liturgical events. Stratigraphic and disarticulation analysis indicates the living returned to Plaza 3A to manipulate the remains not just once but on multiple occasions, sometimes as an appropriate prelude to the killing and dumping of a new group of victims. Sacrifice victim remains required handling and manual disarticulation, despite the wide variety of knives at the disposal of the Moche. It is also worth bearing in mind an iconographic scene of dismemberment that Alva and Donnan (1993: fig. 144) consider as the concluding element of the Sacrifice Ceremony narrative. Here, a pair of female attendants are tugging at the arms of a prisoner and are surrounded by disembodied arms, legs, and a head. Might disarticulation of victims be associated with a gendered division of ritual

9 2   Backo

A

b 3.9. (A) Modified temporal bone from Huaca de la Luna Plaza 3A, where semilunate chop marks were observed superior to the mastoid process (detailed view, inset). (B) Frontal and lateral views of the modified cranium of HG99-5, Huaca de la Luna Plaza 3C. Photos by Heather Backo.

duties, cross-cutting Moche ideas about fertility, sex, sacrifice, and death (Bourget 2006)? One wonders if Moche priestesses held a special role or responsibility in the disarticulation of sacrifice victim remains. A final set of considerations for Plaza 3A involves differences in bone representation (table 3.4) and disarticulation (figure 3.10) between the stratigraphic layers of SS2 and DS2. These two layers were associated with multiple episodes of ritual violence during one of the last periods of heavy rains. Both layers reveal evidence of disarticulation, but the earlier, deeper layer (DS2) is composed of far more complete and intact skeletons. This feature may reflect the fact that DS2 was formed primarily during a flooding event that washed a large amount of sediment from the adobe walls into the plaza, creating a lake of fine mud that solidified around the victims. If the individuals had not yet been disarticulated, the mud would have effectively glued the remains down as they dried. SS2, however, was formed primarily after the rainfall event and is associated mostly with fine windblown sand that facilitated far easier manipulation of the remains (Bourget 1997). A curious pattern of pits present on many of the bones (figure 3.11) correlates to experimental studies that indicate small stones and pebbles in the matrix were pressed into the

The Taphonomy of Ritual Killing  93

huaca de la luna plaza 3a strata ss2 (strata 1)

huaca de la luna plaza 3a strata ds2 (strata 2)

3.10. Disarticulation patterns for the first (SS2) and second (DS2) stratigraphic levels of Plaza 3A. Group 1 (most disarticulated) = medium dark gray, Group 2 = light gray, Group 3 = medium gray, and Group 4 (least disarticulated) = dark gray. Image by Heather Backo.

3.11. Pits present on many of the bones correlate to experimental studies that indicate small stones and pebbles in the matrix were pressed into the bones by pressure from above, possibly by people walking over the remains of previous victims. Photo by Heather Backo.

9 4   Backo

bones by pressure from above (John Verano, personal communication 2008), possibly by the weight of an individual walking over the remains of the previous victims. For all the attention to the complexity of Plaza 3A, Pacatnamú holds a key comparative perspective. This mass grave indeed was far simpler and contained far fewer victims. The patterns of alterations to the bodies and bones in this assemblage represented a very different formation process and interaction between the living and the dead. As with Plaza 3A, the killings at Pacatnamú were accretional, involving an identical victim demographic composition, and bodies were intentionally left exposed for a prolonged interval. Yet other facets of the Pacatnamú mass grave point to significant differences from the earlier Moche. Similar focus on the manipulation and removal of heads is absent, though mandible removal, through some avenue, still apparently held great priority. The physical interaction with the dead at Pacatnamú involved a fundamentally different character and scale. Ritual activities involving targeted element loss, such as seen among the Pacatnamú radii, could reflect two different motivations: ancestor worship or trophy collection. Since the overall picture at Pacatnamú strays so far from the expected treatment of venerated ancestors (Salomon 1995), trophy collection remains the probable motive (Verano 1996, 2001a). The forced disarticulation of elements for trophy collection usually produces, as noted earlier, a characteristic suite of traits in the form of circumferential cut marks, cuts near articulation and tendon attachment points, and damage to articular surfaces (Andrushko et al. 2005; Smith 1997; White 1992). This element of highly targeted trophy collection at Pacatnamú further diverges from the killings at Huaca de la Luna and may ultimately reflect the fact that these deaths involved a more secular motivation involving execution of captives (Verano 1986a; Verano and Phillips, chapter 9). It also highlights a point raised by Klaus and Shimada (chapter 5) that the social and ritual perception of these different victims, the reasons why they were killed, their constructions of personhood, and treatment in death may have differed a great deal between the Moche sacrifices at Huaca de la Luna and later at Pacatnamú. Rather than a shift in religious ethos, the shift within ritual activities and body treatment may reflect differing social identities and social motivations underscoring these two acts of ritual killing. The disarticulation of remains and collection of elements was intended neither to reintegrate the deceased victims at

The Taphonomy of Ritual Killing  95

Pacatnamú into the world of the living nor to ensure that they would “live on” with a new, transformed identity and role.

Conclusion In this chapter, a comparative examination of sacrifice victim taphonomy at Huaca de la Luna and Pacatnamú demonstrated some of the archaeological, analytical, and interpretive challenges associated with highly disturbed and multi-individual depositional contexts. Here, a synthesis of quantitative and qualitative analyses linking estimators of original sample size, element representation, and patterns of disarticulation confidently exclude natural taphonomic agents as the sources of disarticulation, sorting, and element loss in these settings. This novel multimethod approach opens new methodological and interpretive windows on the treatment of the sacrificial dead on the north coast of Peru. Many new questions have arisen in the course of this work to be considered in future investigations. First, to what degree are differences in victim treatment a function of shifting ideas and ideology over time? Alternatively, the immediate situational and contextual associations differed immensely. To what degree did the Huaca de la Luna sacrifices take such forms due to association with acute ecological disturbance (Bourget 1997), versus the Pacatnamú killings, which were likely more secular (but still ritualized) execution of captives or criminals (Verano 2001a, Verano and Phillips, chapter 9)? Also, there should be great curiosity as to why the living chose to manipulate the specific body parts that they did. What were the potential emic meanings of skulls, long bones, and the specific elements targeted for trophy-taking? And while beyond the scope of this chapter, it should be clear that the patterns of bone manipulation in these sacrificial settings demonstrate significant overlap with those seen in north coast non-sacrificial funerary contexts (Gaither et al. 2008; Klaus 2003a, 2008a; Klaus and Tam 2009a, 2015; Millaire 2004; Shimada et al. 2015). The sacrificial setting of Plaza 3A centers on an enclosed, walled area both physically and ritually separated from the other activities that took place beyond the sacred space. The lack of natural taphonomic agents having any influence on formation processes, despite the exposed nature of the decomposing remains, emphasizes the protected nature of the plazas. This is in opposition to standard mortuary practice and space as seen in the well-documented range of Moche burial practices. The liminal

9 6   Backo

state of the decomposing corpses was visually shielded from the populace. However, there would have undoubtedly been other sensory indicators of the activities occurring beyond the sacred walls such as odor and sound (Turner 2011). Despite this, some individuals were not meant to maintain this exclusion. Rather, after their liminal, and perhaps dangerous, state had passed, some may have thus been reincorporated into a traditional mortuary and/or secular setting as grave goods or ritual objects. Further analysis of the potential overlap between mortuary and sacrificial ritual domains is perhaps one of the most important questions yet to be addressed. Implicitly linked to these issues are two methodological points. First, bioarchaeologists should be far more routinely integrated into the field excavation of mortuary and sacrificial deposits in northern Peru, as training in anatomy and bioarchaeology maximizes observations and excavation strategy no matter how straightforward a funerary context may appear. Second, catastrophic or accretional mass burials with a complicated postmortem history are best approached via detailed and multi-method taphonomic approaches—a vital emerging frontier for the mortuary archaeology of the north coast of Peru. Acknowledgments The author would like to express sincere gratitude to the archaeologists and researchers of the Proyecto Arqueológico Huacas del Sol y de la Luna and the Universidad Nacional de Trujillo, Santiago Uceda, Moises Rivero, and Steve Bourget. The Patronato Huacas del Valle de Moche enabled access to the material recovered from Huaca de la Luna, and the Museo de Huaca del Arco Iris extended hospitality and efforts in making the material from Pacatnamú accessible for this research. I would also like to thank Haagen Klaus and J. Marla Toyne for their invitation to participate in the 2011 Society for American Archaeology symposium on human sacrifice, Dr. John Verano of Tulane University for his tireless assistance in reviewing and editing the materials, and the Joint POW/ MIA Accounting Command Central Identification Lab at Joint Base Pearl Harbor–Hickam, Oahu, Hawaii, for their support. The National Science Foundation generously provided the financial assistance for this study. Haagen Klaus, Marla Toyne, and three anonymous reviewers shared helpful editorial comments on earlier drafts of this chapter.

Ritual Strangulation in the Southern Moche World  97

chapter four

Ritual Strangulation in the Southern Moche World Mortuary Ligatures as Tools of Liturgical Violation David Chicoine Department of Geography and Anthropology, Louisiana State University

Since the late twentieth century, scholars have documented elaborate and often spectacular forms of ritual killing and sacrifice throughout the Central Andes (Bourget 1998a, 1998b; Browne et al. 1993; Gaither et al. 2008; Klaus et al. 2010; Murphy et al. 2010; Proulx 1999; Toyne 2008, 2011b; Tung and Knudson 2011; Valdez 2009; Verano 2008a, 2008b; Verano and Toyne 2011; Verano et al. 2007). Various forms of bloodletting, decapitation, chest-opening, heart extraction, dismemberment, and defleshing have been documented bioarchaeologically in association with ritual activities. It is also significant that many complex Andean societies, such as the Paracas, Nasca, Recuay, and Wari, depicted violence and sacrificial practices through visual art and material culture (Frame 2001; Lau 2004; Proulx 2001; Tung 2012). This is particularly the case for the Moche (AD 100–800). On the north coast of Peru, the Moche developed a complex and powerful religious ideology partially enacted through spectacular ceremonies that included myriad depictions of human sacrifice (Arsenault 1988; Bawden 1995, 1996; Hocquenghem 2008; Quilter 2002; Swenson 2003). Excavations at important ceremonial sites such as Huaca de la Luna, Sipán, Huaca de la Cruz, Dos Cabezas, and San José de Moro have yielded material and osteological evidence that substantiate observations made on the basis

9 8   Chicoine

of Moche visual arts and help to explore ritual killing during the first millennium AD (Klaus and Toyne, chapter 1). However, while many of the more sanguineous practices observed in Moche fineline paintings have been observed bioarchaeologically, little emphasis has been placed on Moche ritual killing via strangulation. The goal of this chapter is to explore the settings, meanings, and physical evidence for Moche ritual strangulation in light of recently discovered Moche contexts in the Santa and Nepeña valleys on the southern portion of the north coast of Peru. A literature review generates a working model of ritual strangulation in Andean societies. Based on iconographic and bioarchaeological data, death by strangulation is hypothesized as a special and distinct form of ritual killing. I explore what makes it special, from the physical marks it leaves on skeletal remains to the symbolic meanings associated with this type of ritual killing. I argue that sacrificial strangling—which is difficult to identify archaeologically—has been an overlooked category of sacrifice. In light of recently discovered Moche contexts from the Santa and Nepeña valleys, there indeed exists tantalizing evidence of ritual strangulation. The differences between ligature strangulation contexts and supposedly mainstream Moche religious practices (especially blood sacrifices) suggest the development of new canons of ritual killing inspired by—yet different from—what scholars have described so far for the Moche. Indeed, the Santa and Nepeña contexts, with their symbolic inversion materialized in prone body positioning and less violent or spectacular forms of ritual killing, can be used to explore the diversity of mortuary practices during the first millennium AD. The chapter makes the claim that the combination of physical anthropology, visual art, and archaeological contexts of sacrifice increases our chances to detect and understand the nature, roles, and meanings of ritual strangulation.

The Physical Anthropology of Strangulation In order to study ritual strangulation in the archaeological record, we need to consider the act of strangling and the physical traces it leaves. Strangling produces a form of asphyxia caused from the constriction of the neck without suspending the body (Luke 1967; Purdue 2000). There are two primary forms: (1) strangulation by ligature or garroting, and (2) manual strangling or throttling. Both types lead to a sudden and violent compression of the trachea that causes almost immediate

Ritual Strangulation in the Southern Moche World  99

insensibility and death (Rao 2010). Consciousness is typically lost in five to eleven seconds when the carotid and vertebral arteries are compressed and obstructed (Rossen et al. 1943). The pressure needed to obstruct neck vessels varies from 2kg of force for the jugular veins to 10kg to effectively compress the carotid arteries (Püschel et al. 2004). In the case of ligature strangulation, unconsciousness typically occurs between ten and fifteen seconds1 (Turvey 1996). Death may be caused by (1) asphyxia; (2) cerebral anoxia or venous congestion; (3) combined asphyxia and venous congestion; (4) vagal inhibition (i.e., reflex cardiac arrest; Simpson 1949); and in rare cases (5) fracture and/or dislocation of cervical vertebrae (Rao 2010). Considering the rapidity of unconsciousness and the causes of death, strangulation can be considered to be a rather unspectacular type of killing, especially when compared to bloody and corporally destructive acts of throat-slitting, decapitation, or chest-opening. For bioarchaeologists and forensic anthropologists alike, the problem is that strangulation rarely leaves osteological traces. Neck injuries can occur during strangulation, including soft tissue bruising (i.e., manual pressure greater than ligature; see Iserson [1984]) and laryngohyoid trauma (Adelson 1974; Green 1973). According to Polson and Gee (1973), unless considerable force is applied, damage to neck structures is rare. As pointed out by Aufderheide and Rodríguez-Martin: Evidence for manual strangulation in recently deceased bodies is largely limited to fractures of the hyoid bone and calcified thyroid and cricoid cartilages as well as bruised local cervical soft tissues. In skeletonized bodies the cartilaginous fractures can only be detected if they are calcified, so that such a diagnosis is more apt to be made among the aged. Non-manual (rope, rigid items) strangulation may result in cervical vertebra fracture including hairline transverse fractures on or posterior to the superior articular processes of the axis (Aufderheide and RodríguezMartin 1998:29).

The dimensions, materials, and location of ligatures along the neck column should all be considered when positing forensic guidelines for identifying strangulation. The general consensus, however, appears to be that strangulation primarily involves soft tissues and is unlikely to leave physical traces on the skeletal body. When present in the archaeological record, its evidence may often be quite subtle and contextually

1 00  Chicoine

derived. Its evaluation is further aided by the degree of the preservation of soft tissues and the presence of ligatures around the neck. Given that Andean archaeologists are unusually privileged in the availability of visual art to understand the nature of ancient lifeways in their area of research, I have elected to focus on these cultural artifacts as a guide to understanding ritual strangling on the north coast of Peru. Further, recently discovered funerary contexts with ligatured bodies allow for a preliminary consideration of this form of killing.

Ritual Killing and Strangling in the Central Andes Human sacrifice implies some preprogrammed repetitive action laden with special meanings (Benson and Cook 2001); accordingly, difference and variability in modes and contexts of ritual killing must be perceived as distinct forms of symbolic engagement with the human body (Jean and Zammit 2005; Valeri 1985). According to Green (2001:19), a sacrifice implies “giving up something important to an individual or community for a reason perceived to be of greater importance in some manner than what is to be sacrificed.” As emphasized throughout this volume, sacrificial practices can focus on various types of objects and beings. In the case of humans, life was being offered and bodies and body parts were used as dedicatory offerings, retainers, prisoners, and trophy heads, among other mediums of offering (Verano 1995, 2001a). With respect to the physical evidence of Andean ritual killing, they were varied and included decapitation, blunt force cranial trauma, chest-opening, stabbing, and throat-slitting. Here, particular focus falls on ritual strangling. There are significant ethnohistorical and archaeological data to demonstrate strangling in the Central Andes during pre-Colonial times, with chroniclers such as Cieza de León (1967 [1538]), Betanzos (1996 [1557]), and Molina (1943 [1575]) describing strangulation—along with throat-slitting and tearing out of hearts—as a standard mode of Inka sacrifice (but see Reinhard and Ceruti [2010] for a contrasting argument that Inka sacrifice did not involve heart removal). Victims of strangulation were either interred or cremated. This has been partially confirmed archaeologically. Uhle (1991 [1903]) posited that the series of strangled Inka-period women recovered near the Temple of the Sun at Pachacamac was particularly associated with capacocha events, where the integrity of the victim’s body was central to the sacrificial practice. Uhle went on to state that these sacrifices were significantly different in social terms

Ritual Strangulation in the Southern Moche World  101

from more sanguineous ritual killings seen elsewhere. The Pachacamac women had apparently been strangled using cotton ligatures that were still preserved in situ, and the anatomical evidence of strangulation was unambiguous. This included reduced neck width, stretching of the neck skin, and twisting of cervical vertebrae. More recent work at Pachacamac has confirmed the importance of ritual offerings and killings at the site (Eeckhout and Owens 2008; Shimada et al. 2005a). On the north coast of Peru, pre-Inka groups developed varied traditions of sacrifice. Excavations have yielded insights into ritual killing from Cupisnique to late pre-Colonial times (Alva and Donnan 1993; Bourget 1998b; Cordy-Collins 2001a; Klaus et al. 2010; Toyne 2011b; Verano 1986a, 1986b, 1995, 2001b, 2008). One particularly dynamic sacrificial tradition was developed by the Moche between the first and ninth centuries AD. Archaeological evidence from several sacrificial sites indicates that Moche rituals were institutionalized and involved different activities, including ritual combat, prisoner-taking, blood sacrifice, and dismemberment (Bourget 1998b, 2001a, 2001b; Verano 2008a; Verano et al. 2007). Moche priests developed an extensive repertoire of ritual killing that evidently involved different types of victims—from sea lions to deer, camelids, humans, and inanimate objects—and spanned many different forms of lethal practices and actions. For instance, Bourget (1998b, 2001a, 2001b) reports on El Niño Southern Oscillation (ENSO)–associated sacrifices of young to middle-aged male warriors at Huaca de la Luna’s Plaza 3A. Here, some of the victims were clubbed to death as evidenced by cranial trauma, while many others had their throats fatally lacerated (Hamilton, chapter 2; Verano 1998). Bodies were left exposed and significantly altered by the living. Rotting bodies in many cases seem to have been manually disarticulated, and body parts were often symbolically arranged, according to Bourget (2001b), to emphasize duality, inversion, and the metaphorical and symbolic interplay between the ritual killing of sea lions and human victims. Moche sacrificial contexts—especially the Sacrifice Ceremony iconography—emphasized bloodletting through throat-slitting and the collecting of blood in special semispherical cups (Bourget 2001b; Castillo 2000; Donnan 1978; Hocquenghem 1980). Here, artists depicted blood sacrifice of war captives or ritual combatants that culminated in the presentation of a goblet to a figure known as the Warrior-Priest (Donnan 1988). However, more variants of human sacrifices were practiced

1 02   Chicoine

that have yet to be systematically observed in Moche iconography. For instance, Verano (2008a) and Hamilton (chapter 2) reported on complex victim defleshing practices at Huaca de la Luna’s Plaza 3C that were not depicted by Moche artisans.2 Similarly, until now, there has been little evidence that less bloody methods might have been employed. I suggest that Moche art historians and archaeologists have overlooked practices of ritual strangulation—partially due to the difficulty of recognizing it in human remains, and partly due to the emphasis on other forms of more recognizable ritual killings. The Art of Ritual Strangulation in Moche Visual Culture Moche art is undoubtedly one the most vibrant visual expressions of the ancient Andes. Artwork took many forms and was expressed through various media from body art to murals, textiles, and ceramic vessels. Representations were organized within a coherent system of symbolic communication and expression centered on key aspects of Moche cosmology, beliefs, and ritual life (Bourget 2005, 2006; Donnan 1978; Jackson 2008; Quilter 1997). Complex iconographic scenes and narratives were used to depict a limited number of themes closely related to the liturgical activities of Moche elites, including priests, warriors, temple attendants, and others. Of particular importance for this chapter, Moche artists frequently depicted ligature usage in their painted scenes of ritual killing, where they are used to bind, immobilize, and potentially strangle war captives and sacrificial victims. Captives were stripped of their weapons and other military attire, and their hands, feet, and neck were bound with ropes. They were frequently depicted as being attached to one another with ropes around their necks, forming a human chain. Evidence of binding practices can be clearly seen in the often-cited Sacrifice Ceremony scene (see Klaus and Toyne, chapter 1: figure 1.7), which includes a seated prisoner with bound hands, feet, and neck. It is perhaps significant that ropes were also used to decorate or adorn surrogate sacrificial victims, including the necks of ceramic jars and even deer (Donnan 1982; Hocquenghem 1983). In fact, ceramic vessels were often themselves painted with lines and other rope-like features around their necks and shoulders. In one depiction of a funerary structure or charnel house, ligatured jar offerings are placed in front of the building, perhaps suggesting a link between mortuary rituals and the use of neck ligatures. As pointed out

Ritual Strangulation in the Southern Moche World  103

especially by Millaire (chapter 13) and Klaus and Toyne (chapter 1), inanimate material offerings were actively integrated in sacrificial practices as symbolic forms of ritual killings on the north coast of Peru. The intentionally smashed, unfired clay vessels depicting nude captives in Huaca de la Luna’s Plaza 3A represent a good example of such acts (Bourget 2001b). It is also significant that ropes and ligatures are key elements of the Burial Ceremony that emerges in Late Moche period iconography (Donnan and McClelland 1979; Hill 1998). In their analysis of the Burial Theme, Donnan and McClelland (1979; but see Bourget [2006] for a different interpretation) identify a series of resemblances between Late Moche images of burial activities and firsthand observations made by Antonio de la Calancha in the early sixteenth century. Calancha witnessed a curer beaten and stoned to death as punishment for his patient having died. In another case, a curer was tied with a rope to his dead patient buried underground. Scavenging birds would then eat his body; this was perhaps an attempt to deny the soul a proper and safe transition into the afterlife. Donnan and McClelland (1979:11) point out that every single Late Moche burial scene—interpreted as a ritual of high social and symbolic values—depicted human sacrifices, particularly in the form of nude and splayed female bodies being pecked at by birds in the upper left portion of the scene. Whether or not these particular victims were curers, attendants to elite individuals, or something else entirely is unclear, but the fact that these works depict sacrificial victims offered to birds suggests a Moche habit of ritual killings involving cordage to bind and immobilize victims. Indeed, Middle Moche artists also depicted nude males tied to sacrificial racks with vultures pecking at their eyes (Bourget 2006:202, fig. 4.19). Ropes used to restrain or strangle are often portrayed ending in a snake head, thus resembling the double-headed snake belt worn by the Warrior-Priest and the principal Moche deity that Benson (2012:61–72) designates as the Snake-Belt God. Sacrificial attendants bear semilunar tumi knives with a fox-serpent emanating from the handle, while others hold cups, most likely to collect blood. The focus of this particular ritual setting is so sanguineous that it is difficult to connect ligatures with death by strangulation, since cutting and bleeding was more visibly prioritized. However, there may be a more cogent relationship between the actions of the different modes of ritual killings and the behaviors of predatory animals.

1 04   Chicoine

In the ancient Americas, acts of sacrifice are often symbolically associated with animal predation (e.g., felines, [Saunders 1988]). On the north coast of Peru, Formative Period Cupisnique artists are known for depicting arachnid-like creatures decapitating victims and carrying heads in netted bags (Cordy-Collins 1992). The Moche were keen observers of animal predation, creating many related analogies with ritual killing (Alva Meneses 2008b; Bourget 1994). Their artists often portrayed ritual killers as composite therianthropic beings that combined human and animal attributes. A particularly poignant Middle Moche bottle depicts an anthropomorphized feline captor holding a seated nude and bound male supplicant looking skyward with a fearful expression (Donnan 1978:169, fig. 246). I suggest that the Moche may have symbolically considered the act of strangling as analogous to the predatory behavior of constrictor serpents. South American constrictor snake species most notably include the anaconda (Boa anaconda), which is indigenous to the Amazonian lowlands. In northern Peru, macanche boas (Boa constrictor ortonii)—recognizable by the hexagonal black and yellow markings and shiny black bifid tongue—inhabit the forested areas of coastal desert valleys (Elera 1998:65). Moche art frequently depicts snakes—as portrayed in their undulating bodies and bifurcated tongues—in close proximity to scenes of ritual killing. For example, the Sacrifice Ceremony depicts a rope displaying fox-serpent traits. In a scene derived from the bloodletting sacrificial rite (figure 4.1), a warrior—not yet stripped of his weaponry and military

4.1. Scene of prisoner-taking and bloodletting showing a warrior with ligatures around his neck. Redrawn by Haagen Klaus from Donnan and McClelland (1999: figure 4.103).

Ritual Strangulation in the Southern Moche World  105

4.2. Painted reproduction of the now-destroyed painted murals from the ceremonial center of Pañamarca, Nepeña valley, showing a reproduction of the Moche Sacrifice Ceremony. Painting by Felix Caycho. Courtesy of Lisa Trever and used with permission of Bruna Bonavia-Fisher.

garments—is portrayed with snake ligatures around his neck. The main Moche ritual actor is drawing blood from the victim while one of his attendants is holding the rope with one hand, suggesting indeed a close connection between ritual violence and the use of ligature. A further allusion to the constricting nature of snakes and ligatures can be seen in the snake belt typically worn by the main sacrificer, perhaps implying that the liturgist is taming, appropriating, or embodying the killing power of ropes and other sacrificial tools by wearing them. Indeed, the Pañamarca murals in Nepeña depict a reinterpretation of the Sacrifice Ceremony in which one of the attendants has a rope around his neck, with the rope ending in a semicircular shape reminiscent of tumi knives and Moche warrior backflaps (figure 4.2). If we assume that the seemingly widespread concept of binding and ligature use is linked to the concept of power and control, their presence in the Copulation Ceremony may also be significant (Bourget 2006:162–169). Here, the lower sexual actor, who is sometimes seized by the jaw, has a snake-fox–headed rope tied around his/her neck. In another instance, the penetrator is holding an ulluchu fruit, perhaps emphasizing the control over blood flow, including menstrual and sacrificial blood (but see McClelland 2008) (figure 4.3). In another scene of copulation, a male figure is holding a nude prisoner by the jaw—a clear act of control—while the victim has ligatures around the neck (Bourget

1 06   Chicoine

4.3. Scene of ritual copulation involving the main Moche ritual actor who is wearing a snake belt. Note the ligatures around the neck of the bottom sexual actor. Redrawn by Haagen Klaus from Donnan and McClelland (1999: figure 4.95).

2006: figure 2.10). From what information we have, ritual strangling appears to not just represent a one-dimensional mode of ritual killing; for the Moche, it probably also articulated with attempts at controlling supernatural, natural, or even sociohistorical forces.

The Southern Moche World Moche society is often seen as the first example of state-level political organization on the north coast of Peru and includes extensive material evidence for marked social differentiation (Chapdelaine 1998, 2001, 2011; Shimada 1994a). The sociohistorical circumstances surrounding the emergence and developments of Moche influence are much debated, although it is generally agreed that their visual art and iconography were intimately linked to religious ideology and ritual life (Bourget 2006; Hocquenghem 1987; Jackson 2008; Quilter 2001; Uceda 2010). While the mechanisms of Moche religio-political expansion are still unclear, traditional hegemonic views have been gradually replaced by models that put more emphasis on the heterogeneity and plurality of the Moche cultural sphere (Castillo and Quilter 2010). Scholars now posit a constellation of Moche worlds from Piura in the north to Ancash in the south. In this diverse geopolitical landscape, shared religious practices and identities were visible through ritual performances, including mortuary

Ritual Strangulation in the Southern Moche World  107

patterns and sacrificial practices (Benson 2012; Castillo 2010; Donnan 2010). Recent research has helped to gain a better understanding of historical and cultural complexities in the Southern Moche world. For instance, it now appears that the Gallinazo or Virú styles of ceramics—traditionally considered to be ancestral to the Moche—were in fact sometimes contemporary, suggesting the existence of significant sociopolitical diversity (Millaire and Morlion 2009). At the same time, styles traditionally viewed as Moche (such as Huancaco in the Virú valley) have been shown to represent independent non-Moche polities (Bourget 2003, 2010). Virú indicates the existence of a Moche ceremonial center (Huaca de la Cruz) in close proximity to a local palace complex at Huancaco and coeval Gallinazo centers at Grupo Gallinazo and Huaca Santa Clara (Millaire 2010a). Most of what we know about Moche mortuary rituals is based on research carried out in the northern north coast. However, burials in the southern Moche realm followed very similar, if not shared, ritual grammars with those to the north (i.e., Donnan 1973; Donnan and Mackey 1978; Millaire 2002). The deceased were typically interred underground, in an extended position, on their backs. Corpses were wrapped in textile shrouds and sometimes placed inside a coffin (Donnan 1995:123). Bodies were placed inside chambers that ranged from simple pits to elaborate subterranean tombs. Higher-status contexts tend to be located inside or near monumental buildings, while less elaborate ones are typically found in cemeteries or residential architecture on the periphery of important buildings or centers (Donnan 1995:142, 153; Millaire 2002:169). The Moche seem to have buried many of their dead according to their status and roles in life, such as the priestesses of San José de Moro (Donnan and Castillo 1992; Castillo and Rengifo 2008a, 2008b). Sustained fieldwork at the Santa Valley sites of El Castillo and Guadalupito (Pampa de los Incas) has also brought significant insights into the mechanisms of Moche expansion (Chapdelaine 2008, 2010, 2011). Chapdelaine’s research suggests a gradual arrival of Moche influence through diplomatic alliances with local Gallinazo elites based at the administrative and religious center of El Castillo on the southern margin of the Santa River. The Moche occupation at El Castillo appears to have begun during the fourth century AD in association with Moche III–style ceramics. Chapdelaine (2010:261) hypothesizes that the construction of a stepped platform—bearing representations of Moche-style

1 08   Chicoine

war clubs—on the North Terrace at El Castillo reflects a deterioration of relationships between local elites and invading Moche groups. Moche religion was likely proselytized and negotiated on various bases, including the success of local Moche-related practitioners and the capacity of local communities to negotiate the terms of Moche settlement. Power shifted to the northern margin (Guadalupito) at some point during the sixth century AD, following the abandonment of the El Castillo site and the demise of Gallinazo-style material culture. Guadalupito’s construction is believed to have been associated with a major crisis in the Chicama and Moche regions to the north (Chapdelaine 2000, 2011). Chapdelaine (2010:271) suggests that this coincided with the influx of Moche settlers into the Lacramarca valley from the Moche region along with major modification and extension of irrigation networks and cultivated areas.

Possible Cases of Ritual Strangulation in the Santa and Nepeña Valleys Excavations at El Castillo have yielded human remains in association with the occupation of the summit of the main platform (Sector Alto) as well as the East Terrace (Chapdelaine et al. 2005). The skeletons have been interpreted as pertaining to Gallinazo peoples and reflecting sacrificial events (Gagné 2009:207). Although their exact chronological placement remains unclear, the events likely fall between the third and fifth centuries AD. Eighteen individuals were recovered, six of which were too poorly preserved to be analyzed. Osteological analysis of the remains by Gagné (2009) indicates that half of the analyzed individuals show signs of perimortem trauma, including mutilation, cranial blunt force trauma, and strangulation. Several of the mortuary contexts are notable for containing individuals interred in a prone position, contrasting with typical patterns observed on the north coast during the Early Intermediate Period (Donnan 1995; Donnan and Mackey 1978; Millaire 2002; Tello and Delabarde 2007). A 14- to 16-year-old male (Burial 8) was found with a rope around his neck. A 25- to 30-year-old male was also found in a prone position, with his ankles tied. His skull and two cervical vertebrae were missing. Cut marks suggest decapitation and the perimortem dismemberment of the left arm. The El Castillo victims display parallels with a series of graves excavated at the site of Huambacho, in the Nepeña valley, some 50km to the south (Chicoine 2011).

Ritual Strangulation in the Southern Moche World  109

The Nepeña valley is known for housing the southernmost ceremonial complex with Moche-style paintings and murals at Pañamarca (Bonavia 1985; Schaedel 1951; Trever et al. 2013). However, burial contexts were poorly documented until excavations at Huambacho in 2003 and 2004 revealed the first undisturbed interments in the valley (Chicoine 2011). The situation in Nepeña during the Early Intermediate Period is still insufficiently understood, although it would appear that the Moche presence was limited beyond the vicinity of Pañamarca (Chicoine 2004; Proulx 1982, 2004). Huambacho was mainly built during the Early Horizon (800–200 BC) and re-occupied as a cemetery during the Early Intermediate Period (AD 1–800) by groups burying their dead with Gallinazo or Virú and Moche-style ceramics (Chicoine 2006, 2011). Graves with Late Intermediate Period (AD 1000–1470) Chimú-style vessels were also excavated. At least 23 individuals were documented, buried in simple pits that truncated Early Horizon deposits: further details of excavation procedures and the burial contexts are presented elsewhere (Chicoine 2011). A clear distinction in body position is observable between the Gallinazo/ Virú and Moche graves: the Moche interments are extended and prone, whereas all other burials are flexed. The prone position is unusual and has been interpreted as inversion of the Moche standards of appropriateness associated with contexts of funerary and dedicatory offerings (Hill 2003:287; Chicoine 2011:543). The Huambacho Graves The current study focuses upon three of the 23 burial contexts containing individuals associated with ligatures and showing possible evidence of death by strangulation (Graves 10, 11, and 15). The graves were found at Huaca-A, a small mound located to the northeast of Huambacho’s Main Compound (figure 4.4). This area has been interpreted as a small cemetery used by surrounding local communities during the Early Intermediate Period. The general simplicity of the burial pits, the absence of coffin remains, the remoteness from residential areas, and the modest quality and quantity of grave goods point toward a cemetery for non-elite individuals. Although the elaboration and overall organization of the graves align well with lower-status burials from Pacatnamú and Huacas de Moche to the north (Donnan and Cock 1997; Donnan and Mackey 1978; Verano 1997a), the Huambacho burials stand out for

1 1 0  Chicoine

4.4. Map of Huambacho, Nepeña valley, showing the location of the graves excavated in 2003 and 2004. Map/photo by David Chicoine.

the occupants’ prone position, hinting at a symbolic inversion (Bourget 2006) perhaps linked to the special or deviant nature of the contexts (Murphy 2008). The graves are also notable for being the first undisturbed contexts with Moche-style artifacts to be found in Nepeña. All of the skeletons were oriented toward the southwest and thus aligned with Early Horizon structures (N41oE). Grave 10 Grave 10 is a simple pit some 50cm deep containing the remains of an adult male aged between 25 and 30 years. The individual, who had been wrapped in a cotton shroud, was well preserved naturally with mummified hands and feet. He was wearing a cylindrical headdress made of plant fibers, and a textile bag was placed on the back of his left shoulder. The bag contained dried coca leaves and a bone spatula and resembled bags depicted in the fineline Coca Ceremony scenes (see Chicoine 2011:537, fig. 8). The adult individual in Grave 10 displayed a healed fracture of the proximal diaphysis of the left ulna as well as a healed fracture to the left ninth rib. No ligatures were recovered during excavation.

Ritual Strangulation in the Southern Moche World  111

Grave 11 Grave 11 was located roughly a meter to the west of Grave 10, some 40cm deeper, and contained three individuals: two adults (Individuals 1 and 3), and an infant (Individual 2) (figure 4.5). The two adults are interpreted as part of a single, related funerary phase, whereas the infant is likely to have been buried later, perhaps in Moche times, as part of an intrusive event. Individual 1 was a male aged between 25 and 30 years at the time of death. His hands were placed in front of the pelvis. The bones of the lower legs had been removed after burial and replaced with the skeleton of Individual 3—an adult male aged between 30 and 35 years old. Individual 3 was found in a similar position as Individual 1 but was accompanied by more extensive grave goods (see below), suggesting elevated status. Ligatures made of twisted reed materials were found around the neck, hips, and knees (figure 4.6). The midbody ligatures tied the hands to the hips, while those around the neck suggest strangulation. The cervical vertebrae show no signs of trauma. Two canchero-shaped gourd containers were placed laterally at the thoracic level. The handles of the gourd cancheros point in opposite directions. An additional spherical gourd container was placed near the left hand. Various large sticks were also recovered. The right hand was placed in front of the pelvis underneath the body, while the left was folded behind the back. Copper was found in both hands, as well as in the mouth. Two small copper disks—possibly from hair tweezers— were found near the left shoulder, wrapped in a cotton cloth. A needle and stone with cotton yarns were placed on the back of the individual. Two ceramic vessels were found near the head: (1) a lateral handle bottle in the form of a condor (Vultur gryphus), and (2) a neck jar with a lateral handle in the shape of a toad (Bufo sp.). The deceased was wearing a relatively simple necklace made of carved nacre and Spondylus shell beads. Individual 3 displayed an antemortem fracture in the form of a healed compression fracture to the body of the sixth cervical vertebra, resulting in a localized kyphosis of the neck. The C7 vertebra displays anterior bony outgrowths on the centrum, which formed to compensate for the compression fracture. Two healed rib fractures (right ninth rib, left eleventh rib) were also observed.

1 1 2   Chicoine

A

b

c 4.5. Photographs and drawing of Grave 11 showing the position of the skeletons and grave goods, including (A) an overall view of the context, (B) a detailed view of Individual 3, and (C) remains of one of the gourd grave goods. Map/photo by David Chicoine.

Ritual Strangulation in the Southern Moche World  113

4.6. Photographs of the preserved ligatures tying Individual 3’s hands to his pelvis (Grave 11). Map/photo by David Chicoine.

Grave 15 Grave 15 was found 175cm below the surface, directly beneath Grave 11. Based upon their location, orientation, and stylistic/symbolic affinities, it is likely that Graves 11 and 15 represent historically related episodes, most likely within the range of generational memory. Grave 15 contained the remains of an adult male aged between 25 and 30 years. The body was wrapped in a plain weave cotton shroud and interred in an extended and prone position (figure 4.7). The head was oriented toward the southwest and slightly turned to the left (facing east). In contrast to other skeletons, the bones were clean of soft tissues and lightly sun-bleached, suggesting the body was exposed in a state of advanced decomposition before final burial. The position was also unlike that of the other individuals, possibly suggesting special treatment. The torso was turned to the left, with the body partly resting on its right side. The feet, meanwhile, were turned to the right (west), giving the body a contorted appearance. The neck of the individual was tied to a wooden post, and the hands and feet were bound. The ligature tying the head to the post was made

1 14   Chicoine

A

b

c

4.7. Photographs and drawing of Grave 15 showing the position of the skeleton, grave goods, and knot tying the neck to the post, including (A) a view of the compressed superior portion of the skeleton, (B) placement of grave goods and ligature tied around the neck tethered to a wooden post, and (C) a detailed view of the neck ligature, a copper object in the mouth, and shell pendant. Map/photo by David Chicoine.

Ritual Strangulation in the Southern Moche World  115

of cotton, thus contrasting with the hand and feet ligatures, which were made of coarser reed materials. The neck ligature was tied in a sliding knot at the back of the neck. While this suggests strangulation, the cervical vertebrae displayed no signs of trauma. The adult in Grave 15 displays signs of a physically active and strenuous lifestyle, including widespread degenerative joint disease. Several grave goods accompanied the body, including a copper plate in the mouth, sheets of copper in the right hand and by the neck and right foot, a Moche-style neck jar sculpted in the form of a toad, a jet mirror, a spherical gourd container, a stone polisher, and a shell pendant. The presence of grave goods in association with what appears to be a potential case of ritual killing is puzzling and could point toward the unusual or unique symbolic meanings associated with ligature strangulation.

Discussion The data presented in this chapter comprise the first detailed consideration of rituals of sacrificial strangulation in Moche times, especially on the southern north coast of Peru. A review of Moche visual arts and iconography suggests that ligatures were used in the context of mortuary rituals, as well as in prisoner-taking, the ritual hunting of deer, and the sacrifice of objects and human captives. In the Burial Ceremony, ligatures are used to tie an individual to the ground in what appears to be a torture and/or excarnation ritual, as well as for lowering bodies and coffins into the ground. At the more symbolic and metaphorical level, one might argue that the act of binding and restraining with ligatures involves concepts of control. It is perhaps significant that sacrificial practices involving ligatures contrast with the more ostentatious blood sacrifices during which human bodies become arenas for spectacular display (Hill 2003, 2005). In the case of ligature strangulation, one can suggest a different intention. One might surmise that this assumed some level of symbolic significance: blood sacrifice—in particular throat-slitting— could reflect the predatory behavior of felines, while strangulation may be related to the constrictor properties of some predatory snakes. It is difficult, of course, to demonstrate such a link with certainty, although it may be significant that the main ritual killer in Moche iconography typically wears a snake belt. Whatever the significance, the iconographic evidence certainly points toward an inversion of the typical sacrificial values of spectacle (e.g., bloodletting, mutilation, dismemberment),

1 1 6   Chicoine

strangling being a much more restrained, controlled, and quiet form of killing. Such non-modal practices are often associated with so-called deviant burials and other forms of atypical mortuary treatment (e.g., apotropaic practices that aimed to deflect harmful influences [Gell 1998:83]) (Murphy 2008; Saxe 1970). The concept of deviant burial was developed in the 1930s with the objective of explaining the discovery of burials that contrasted with the modal funerary behaviors of a given society or time period. German-speaking scholars developed the term Sonderbestattung to refer to “special” or “exceptional” burial (Wilke 1933), while in the English-speaking literature the concept of “deviant” burial was adopted (Aspöck 2008; Saxe 1970). In both traditions, focus was placed on the particular status of the deceased as well as the extraordinary circumstances of one’s death (Aspöck 2008). As pointed out by Saxe (1970:225–226), the social persona and the circumstances surrounding one’s death can impact the obligations of those in charge of mortuary rituals. At the negative end of the spectrum, social outcasts, scapegoats, people of low rank, and the ill or disabled can find themselves targets of ritual killing and other forms of mortuary violation. Deaths perceived as potentially dangerous—such as executions, suicides, and early death (mors immatura)—can equally justify an inversion of funerary conventions and a denial of proper mortuary treatment. Here, violations can aim at denying the deceased a proper resting place and prohibiting the adequate movement of the soul through the different stages of the afterlife (Bloch and Parry 1982). At the other end of the spectrum, mortuary contexts associated with high-status religious and political offices, as well as particularly desirable deaths, have the potential to justify atypical mortuary practices (Aspöck 2008:22; Meyer-Orlac 1997:10). As pointed out by Bloch (1992:228), practices of mortuary violation and veneration are part of a dialectic ideological program that can be contested and negotiated over time and space. As paraphrased by Duncan (2005:10), “some acts of violation may create symbols from the deceased’s body parts that are treated reverentially in later rituals.” The sociohistorical circumstances of the arrival of Moche religious influence in the Santa and Nepeña regions could have provided a laboratory for the negotiation and transformation of patterns of social acceptance and associated mortuary treatment and ritual practice. The contexts discussed in this chapter contrast so markedly from typical patterns of Moche blood sacrifice and funerary interment that it might be useful to discuss them from the perspective of deviant, atypical,

Ritual Strangulation in the Southern Moche World  117

or non-modal mortuary practices. Typically, the killing of war captives emphasizes torture, as well as the dismemberment, manipulation, and display of body parts (Verano and Phillips, chapter 9). While the graves from El Castillo display evidence of intense bodily violations (i.e., mutilation, blunt force trauma), the use of ligatures at Huambacho suggests a different form of violation—one that was less intrusive and not centered on the destruction of the body (Hill 2005). Rather, the integrity of the deceased’s bodies and the inclusion of grave goods suggest different symbolic meanings. The Huambacho graves are particularly puzzling for combining mortuary elements that bear both negative and positive connotations. The prone arrangement of bodies and use of ligatures to restrain or strangle the deceased point to intentional negation of proper burial treatment. Yet the evidence does not align well with more extreme forms of negative predation often materialized in the defacement, decapitation, and bodily obliteration of victims (Duncan 2005:211). As pointed out by Bloch (1992), acts of mortuary veneration can also use violence as a liturgical mechanism, especially in contexts in which it is directed at forces that threaten society rather than the individuals’ own persona. Here, mortuary contexts and their associated rituals have the potential to become indexes of social memory and veneration by members of the community. At Huambacho, it is perhaps significant that Grave 15 displays the strongest case for death by strangulation. The victim’s body was likely exposed for some time, as indicated by the surprising lack of soft tissues and the mild bleaching of bones, suggesting that the strangling was followed by excarnation. Within generational memory, two other adult males (Grave 11) were buried directly above the man in Grave 15. One of them was tightly bound and accompanied by relatively elaborate grave goods, while the other had the bones of his lower legs removed postmortem. The complex grave arrangement parallels the complexity of the ritual interments and hints at various moments of grave reopening. The continuity in prone body placement over time, as seen in Grave 10 and the other Moche phase contexts excavated at Huambacho (Chicoine 2011), suggests the negotiation and incorporation of this atypical practice into local funerary programs. Yet it could well have originated in the ritual strangling of undesirable individuals at a time of heightened social and religious tensions associated with the arrival of Moche influence in the southern portion of the north coast.

1 1 8   Chicoine

On the basis of the current limited evidence from the southern Moche world, I would suggest that this inversion might reflect a form of social censure that also included interment in a prone position (Murphy 2008). In the specific case of Nepeña, it is possible that strangulation was a way for some groups to materialize sociopolitical and religious differences. Strangulation could have been a way for local liturgists to develop their own style of killing, inspired by the Moche Sacrifice Ceremony yet different and set apart. This assertion remains tentative pending further evidence of ritual strangulation, but the discussion of the Southern Moche mortuary contexts demonstrates the challenges that Andeanists face when trying to disentangle the concepts, actions, and materializations of ritual killing and atypical mortuary treatments, especially strangulation and the use of ligatures.

Conclusion This chapter has brought new insights into an understudied form of ritual killing in the Central Andes. Through a case study from the Moche of the north coast of Peru, I would propose that strangulation is a significant form of ritual killing that differs socially from more spectacular and osteologically visible sacrificial methods. A review of visual art and iconographic narratives suggests that death by strangulation using ligatures represented a special form of killing. In the case of the southern Moche world, I would posit that ritual strangulation was perhaps part of the development of practices inspired by—yet different from—Moche mainstream religious ideology. While these proposals are necessarily tentative at present pending more data (especially from the monumental complex of Pañamarca), it would conform to our current perceptions of political and social conditions linked to the spread of Moche religious and political influence. In any case, the Huambacho data comprise instances of ritual strangling that differ markedly from established strangling cases such as the cemetery of the Chosen Women at Pachacamac, a few other retainers on the north coast such as at the site of El Brujo (Verano 2001a), or inferred or possible cases at Sipán and Middle Sicán elite tombs (Bentley and Klaus, chapter 10). As seen in this chapter, the anatomical and forensic identification of ritual strangling in Moche remains is still tentative. However, iconographic, mortuary, and contextual evidence all suggest that Andeanists need to pay more sustained attention to the recognition and understanding of cases of ligature use and ritual

Ritual Strangulation in the Southern Moche World  119

strangulation. Beyond the problem of identifying ligature strangulation, future research should pay particular attention to the treatment of deviant, ostracized, or marginal groups and the role of torture, submission, and display in shaping ritual violence in ancient Peru. Acknowledgments I would like to extend warm thanks to Haagen Klaus and J. Marla Toyne for their kind invitation to participate in this volume. Special mentions go to Emily Grace, who helped with the osteological analysis of the Huambacho skeletons. Mary Lee Eggart and Jeisen Navarro worked on some of the drawings. Comments from Haagen Klaus, Élise Lassonde, J. Marla Toyne, and three anonymous reviewers helped shape the final version of this chapter.

Notes

1. If the trachea is partially closed, buzzing in the ears, congestion and cyanosis in the head, vertigo, tingling, muscle weakness, bleeding from the mouth, nose, and ears, clenching of the hands, and convulsions may all occur before death (Rao 2010). 2. Some nude and bound individuals in Moche art are tied to a rack or post and may be represented with their faces having been flayed off. Skeletal beings of Moche art—which may be depictions of dead or ancestral entities—are indeed reminiscent of some of these defleshing practices. Some scholars speculate that there were conceptual links between the two.

1 2 0  Klaus and Shimada

chapter five

Bodies and Blood Middle Sicán Human Sacrifice in the Lambayeque Valley Complex (AD 900–1100) Haagen D. Klaus Department of Sociology and Anthropology, George Mason University

Izumi Shimada Department of Anthropology, Southern Illinois University

The Middle Sicán society (also known as the Classic Lambayeque culture) ascended to regional dominance on the north coast of Peru between AD 900–1100 (Shimada 1990, 1995, 2000; 2014a, 2014b, 2014c). They inherited elements of the Moche legacy but also expressed conspicuous discontinuities with their predecessors—including a shift away from narrative artistic expression. Here, we characterize what transpired to regional practices of human sacrifice following Moche times through a cross-contextual, regional, and interdisciplinary synthesis of 37 years of Middle Sicán archaeology within their Lambayeque Valley Complex heartland. Did the comparative scarcity of representations of ritual killing in Sicán visual art reflect a parallel decline of sacrifice in religious practice? Conversely, if sacrifice persisted, what forms did it assume, and was it linked to earlier cultural precedents?

Archaeological Setting: An Overview of the Middle Sicán Culture For much of the first millennium AD, the Moche dominated the cultural landscape of Peru’s north coast. Moche archaeology and art history have had a long and prominent past (Benson 1972; Donnan 1978; Shimada

Bodies and Blood  1 2 1

and Vega-Centeno 2011). Moche art history has long recognized that violence, death, and sacrifice were core elements of their belief system (Donnan 1976). In the sixth century, a severe, prolonged drought and powerful El Niño Southern Oscillation (ENSO) event were associated with significant sociopolitical and ideological turmoil (Shimada et al. 1991; Winsborough et al. 2012). While the southern Moche largely fragmented, northern Late Moche peoples staved off their final political disintegration for another 200 years (Bawden 2001; Castillo 2001; Shimada 1994a; Swenson 2012). Origins and Organizational Characteristics Following the abandonment of the Late Moche urban center at Pampa Grande, a power vacuum enveloped the resource-rich, densely populated, and strategic Lambayeque Valley Complex (consisting of the Motupe, La Leche, Lambayeque-Reque, and Zaña valleys). North Highland Cajamarca, Wari, and even possibly Pachacamac cultures aimed to expand into the region (Montenegro 1993; Shimada 1990, 1994a), but a new local polity—the Sicán—ascended to regional supremacy between AD 850–900 (Shimada 2000, 2014a). The Middle Sicán period (AD 900–1100) encompassed their florescence and represented the apogee of the northern north coast’s autogenous cultural development (Shimada 2014a, 2014b, 2014c). In as little as a hundred years, they developed a theocratic state and economic powerhouse based on long-distance trade and semi-industrial–scale production of metals as they ushered Peru into the Bronze Age. The Middle Sicán political and economic system incorporated an estimated 1.5 million people over a 400-kilometer coastal stretch, from the Piura region to the north to at least the Chicama valley to the south (Franco and Galvez 2005; Prieto 2010; Segura and Shimada 2014). Instead of militaristic conquest, the Sicán leadership evidently employed cooperative and incorporative strategies that provided economic benefits and prestige for participants (Prieto 2010). The physical, political, and spiritual center of Middle Sicán culture was the religious precinct of Sicán1 located on the banks of the La Leche River (figure 5.1). Late Moche monumental truncated pyramid construction techniques were revived, and the leadership erected six enormous adobe-brick platform mounds, five of which still rise above the canopy of the Pomac desert forest (Shimada 2014a). The capital formed a

1 2 2   Klaus and Shimada

5.1. Plan view of the religious precinct of Sicán in the La Leche valley. Map by Izumi Shimada.

planned ritual precinct (ca. 1km north-south and 1.5km east-west) and was a locus for ancestor veneration (Matsumoto 2014; Shimada 2014a; Shimada et al. 2004). The leadership controlled the largest inter-valley irrigation system in the New World and presided over a rich and multifaceted resource base and labor force. Political power extended into secondary monumental satellite centers at Túcume, La Pava, ChotunaChornancap, Illimo, Patapo, Vista Florida, Sipán-Collique, Luya, Huaca Pared-Uriarte, and Úcupe (Alva and Alva 1984; Martínez 1997; Shimada 2000; Wester 2010). Even rural communities such as Huaca del Pueblo Batán Grande and Huaca Sialupe intensely contributed to the economy (Cleland and Shimada 1998; Shimada and Wagner 2007). Social Organization Funerary variation, skeletal biology, settlement patterns, production spheres, artistic representations, and ascribed value of metal and ceramic products independently attest to the hierarchical and multiethnic nature

Bodies and Blood  1 23

of Middle Sicán social organization (Klaus et al., in press; Shimada et al. 2004). A relatively small group of ethnically Sicán lords occupied the apex of this culture. Elite Sicán individuals were granted burial in exclusive, planned cemeteries at Sicán. Their burials were by far the most materially abundant, appointed with the full range and highest quality of metals (high-karat gold alloys, silver, tumbaga alloy [copper, silver, gold, and arsenic], and utilitarian bronze objects) (Shimada 1994c; Shimada et al. 2004). Inherited dental traits and mitochondrial (mt) DNA variation indicates that members of the elite were very closely related (Corruccini and Shimada 2002; Klaus et al., in press). Also, mortuary pattern,

5.2. Huaca Loro Burial 6 HL-T1-95, which illustrates the hallmarks of a typical burial of a Middle Sicán period commoner burial. Drawing by Rafael Vega-Centeno and Cesar Samillán.

1 2 4   Klaus and Shimada

artifactual, and mtDNA data suggest the elite had ties to southern Ecuador (Shimada 1995; Shimada et al. 2005b). Skeletal data also indicate the elite lived lives of privilege, as they experienced very good health, engaged in physically undemanding lifestyles, and ate high-quality diets (Farnum 2002; Klaus et al., in press; Muno 2014). A lower nobility also existed, but very few of its members have ever been found (Alva 1985). Non-elites, or the social commoners, were the laborers, fishers, farmers, potters, and metalsmiths who represented the Middle Sicán society’s productive core. Commoner burials at the capital (figure 5.2) (Shimada 1995), Huaca del Pueblo Batán Grande (Cleland and Shimada 1992), Huaca Sialupe (Klaus 2003a), Illimo (Martínez 1997), Túcume (Narváez 1995a), Jotoro (Martínez 2011), and La Pava (Fernández 2011) follow consistent patterns that include burial with only bronze items or no metal at all; positioning of the deceased in an extended position oriented to a cardinal direction (most often south-north, with the head to the south); placement of no more than ten ceramic vessels in a grave; and a few sets of camelid extremities (Shimada 2000). This represents continuity of a formal ritual grammar that emerged during the Moche period (Millaire 2002). These burial patterns, along with diverse artistic evidence (Shimada 2014a, 2014b), simultaneously indicate memory and expression of conserved Moche-like (or Mochicoid) styles and identities among most of the local peoples. They were apparently members of a non-elite and ethnically Muchik substratum under the surface of Middle Sicán society (Klaus 2014a). Persistence of Moche descendants should not be surprising, as there is no evidence of a population extinction or replacement following Late Moche times. As members of subaltern social strata, these people experienced lives quite distinct from those of their rulers. Higher prevalence of chronic childhood biological stress, osteoarthritis, and lowerquality diets are strongly linked to Middle Sicán–era commoners in both the heartland and provinces (Farnum 2002; Klaus et al., in press). Artistic Representations of Sacrifice Middle Sicán religious art embodied an innovative configuration that hybridized earlier Moche and Wari and Pachacamac traits into a new configuration (Shimada 2000). The hallmark of Middle Sicán corporate art was the Sicán Deity. This iconic being, with its upturned, winged eyes, triangular nose, and stylized pointed ears, dominated expressive media. Linked to the nocturnal raptorial bird of Early Sicán art, it connoted

Bodies and Blood  1 2 5

cosmological symbolisms of omnipotent power related to the moon, sun, and life-giving rains (Shimada 2000, 2014b). Tello (1937) and Carrion (1940) considered the Sicán Deity as one of sacrifice, as it was sometimes depicted holding a crescent-bladed tumi in one hand and a kero-like flaring cup (or, more rarely, a trophy head) in the other hand. The Sicán Deity also was represented in superb detail on gold tumis. Some have speculated such knives were sacrificial implements (Benson 2001:6); more likely, the malleability of the blades made them functionally ill-suited to cutting, relegating them to ceremonial or status items. All currently known Middle Sicán sacrifice-related imagery is observed either in metalwork associated with high-status burials at Sicán, or in textiles attributed to the famed religious center of Pachacamac, particularly those curated in the C. T. Wilhelm Gretzer Collections in the Ethnological Museum in Berlin (Hoffman 2007; Segura and Shimada 2014). A looted Late Middle Sicán silver kero at the Denver Art Museum features complex repoussé decoration (Mackey and Pillsbury 2013). It depicts elements of a remarkable and rare narrative-like cosmovision with interlinking scenes of procession and presentation, a water channel, activities in a coastal landscape, and an avian theme. Populating these themes are hundreds of figures, waves, trees, individuals sprouting plants from their bodies, and two prominent “entombed” females, all presided over by a dozen figures atop the scene. The Denver Kero appears to convey a principal message about death, burial, and regeneration as noted elsewhere in Middle Sicán art (Shimada 2014b). Numerous disembodied heads, trophy heads, and a pair of headless bodies accompanying one of the entombed females in Section B are suggestive that sacrificial rites are depicted on the Denver Kero. Linked and interlaced visual metaphors of liquids and vegetal fertility are rather explicit. A pair of large painted textiles found in the partially looted Huaca Las Ventanas South Tomb depicted the Sicán Deity in the pose of a Moche Decapitator (figure 5.3). On another textile, the Sicán Deity, similarly holding a tumi and a trophy head, is under the night sky represented by an overarching bicephalic serpent. On the other textile, the similarly posed deity was surrounded by zoomorphic ocean waves with the sun on one end (found oriented to the east in the tomb) of the scene and a crescent moon (on the west end) on the other. These may represent a cosmogonic event (creation of the universe, moon, and sun through sacrifice) or a cosmovision representing the Sicán Deity balancing the structure of the universe through the taking of life. Elsewhere, a painted textile in the

1 2 6   Klaus and Shimada

5.3. Reconstructed rendering of the center image of a large painted cloth documented in the partially looted Huaca Las Ventanas South Tomb at the Sicán capital. It depicts the Sicán Deity in the iconic pose of a Moche Decapitator holding a crescent-shaped tumi in one hand and a severed human head in the other and probably represents a cosmovision. Painting by Cesar Samillán.

Huaca Loro West Tomb depicted marching Sicán warriors, each holding a decapitated human head in the right hand (Shimada 2014a). The Gretzer Collections contain looted Middle Sicán–style textiles that depict blood poured into a goblet among other scenes involving concerns over agricultural and fishing success and fertility (Shimada 2014a).

Evidence of Middle Sicán Ritual Killing For most of the twentieth century, the Sicán culture was chronically understudied. In 1978, Izumi Shimada initiated the Sicán Archaeological Project. After establishing Middle Sicán chronology, ecology, and technology, focus turned to the interdisciplinary study of social organization, mortuary practices, and population biology in 1990. Here, we examine data from 483 Middle Sicán individuals recovered by the Sicán Archaeological Project and parallel work by our Peruvian colleagues at various sites in the region. These burials span all known social variation (Farnum 2002; Klaus et al., in press; Muno 2014; Shimada et al. 2004). Our approach examines sacrificial contexts in comparison with this regional sample of habitational, workshop, ceremonial, and funerary settings

Bodies and Blood  1 27

while integrating contextual data on sacrificial preparations, feasting, burial, offering, remembrance, and related activities. Thus, our basic approach is cross-contextual, regional, diachronic, and interdisciplinary as advocated by Buikstra (1977; also Klaus 2008a; Shimada et al. 2004). Dedicatory Burials and Tomb Retainers at the Sicán Capital An essential focus of Middle Sicán religion appears to have involved the veneration of divine elite ancestor lineages (Matsumoto 2014; Shimada et al. 2004; Shimada 2014b). At Sicán, we hypothesize that the six primary adobe platforms (Huacas Loro, Lercanlech, Las Ventanas, El Corte, La Merced, and Sontillo) each represented a physical locus for worshiping a particular lineage (Shimada 2009, 2014a). Large roofed altars atop these adobe mounds functioned as ancestor veneration temples and were perhaps the most sacred spaces in the entire Sicán world. The first category of suspected sacrificial contexts—dedicatory burials—are associated with these temples. Large, consistently spaced wooden posts (3 meters apart on both the east-west and north-south dimensions) were set into deep subfloor sockets or boxes to support the temple roofs. Erosion has severely damaged the remains of each temple, but excavation of surviving floors atop Huacas Las Ventanas, Lercanlech, Loro, and Sontillo indicate a highly repetitive pattern where every other columnar socket contained a human body. Alternating sockets contained bundles of standardized bronze sheets often paired with a whole Spondylus princeps shell (Bezúr 2003, 2014; Shimada 1990, 1995). A subfloor columnar box burial atop Huaca Loro’s North Platform contained the body of a seated old adult male. He wore a woven cotton blindfold, and his arms and legs embraced the post in front of him (figure 5.4a) (Muno 2014). Identical contexts were documented atop Huacas Lercanlech (figure 5.4b) and Las Ventanas. In another similar case at Huaca Las Ventanas, a man was placed on his back amid the construction of the central access ramp (figure 5.5). His chin was neatly tucked against his chest, his arms were crossed at the waist, and his legs were drawn up and flexed. This individual was also blindfolded, and his wrists were bound with cotton cordage. The second kind of inferred sacrifice was associated with elite funerary settings. Principal personages were accompanied by the bodies of retainers or family members, as documented in the Huaca Loro East, West, and Northeast tombs, the West Cemetery, and the tomb of a lower

1 2 8   Klaus and Shimada

A

b

5.4. Two examples of columnar box burials atop (A) Huaca Loro and (B)Huaca Lercanlech. Photos by Izumi Shimada.

5.5. A bound and blindfolded man placed into the central access ramp of Huaca Las Ventanas. The left leg was disturbed and removed in antiquity. Photo by Izumi Shimada.

elite in nearby Illimo (Farnum 2002; Klaus 2008a; Martínez 1997; Shimada 1995, 2014a; Shimada et al. 2000, 2004). Bioarchaeological study reveals the majority of tomb retainers were nearly always young and middle-aged adult women. These retainers were skeletally healthy at the time of death and lacked signs of recent illness or any kind of perimortem traumatic injury (Farnum 2002; Muno 2014).

Bodies and Blood  1 2 9

Some retainers in the Huaca Loro West Tomb (figure 5.6) (Shimada et al. 2004) were missing terminal finger and toe phalanges. In others, skeletal elements were disarticulated in ways inconsistent with post-depositional decay of wrappings, soft tissue decomposition, or settling of grave fill (Farnum 2002; Shimada et al. 2004). For example, the entire right arm of West Tomb Burial 7 was found under the left side of the vertebral column. A few empty fly pupae (Faulkner 1986; Huchet and Greenberg 2010) were documented in direct association with Burials 9 and 14. Combination of disarticulated bones, missing phalanges, and necrophagous insects is convincing evidence of prolonged curation of their bodies (Shimada et al. 2015). Some Middle Sicán retainers evidently died months, years, or even generations before (Bentley and Klaus, chapter 10). Considering the scale and complexity of elite funerals, preparation almost certainly took place while a Sicán lord was still alive and may have included curation of already-dead individuals. In Huaca Loro West Cemetery, an adult woman with a detached skull and a headless adult man (figure 5.7) were found one atop the other adjacent to the mouth of Tomb 1, which contained a young noblewoman. Their positioning and stratigraphic positions suggest they were tossed in and were soon thereafter overlain by a thick flood deposit. These unique burials occurred close to, or at the onset of, a major ENSOrelated flood (see below) conceivably to propitiate an ancestor below. Another unique context (HLV Burial 7) was found just beyond the eastern façade of Platform 1 at Huaca Las Ventanas. Here, the head of a 15- to 20-year-old male was found resting on his feet. Cut marks on the first two cervical vertebrae (still attached to the base of the skull) indicate that his throat had been deeply lacerated, he was decapitated, and the head placed upon the feet, perhaps in a symbolic inversion of normal anatomical form (Klaus 2009). So far, this is the only known Middle Sicán case of unambiguous decapitation. This case could represent a more secular punishment (sensu Verano 1986a) but could be something akin to a so-called foundation sacrifice as well. Ritual Killing Beyond the Capital Outside of Sicán, sacrifice had a different character. As with mortuary rituals, sacrifice appears to have persisted as a deeply rooted, traditional practice of the non-elite Muchik ethnic peoples. At the multi-craft workshop of Huaca Sialupe in the lower La Leche drainage, the burial

1 3 0  Klaus and Shimada

5.6. Overall layout of Huaca Loro West Tomb (top) with a detail of one pair of young female retainers (bottom—Burials 11 and 12) crammed into Pit 5 on the antechamber floor. Drawings by Izumi Shimada and Cesar Samillán.

Bodies and Blood  1 31

5.7. One of the two headless adults found next to the mouth of Huaca Loro West Cemetery Tomb 1, a shaft tomb that contained a young adult noblewoman. Their death and burial appears to have occurred close to or at the onset of a major ENSO-related flood. Photo by Izumi Shimada.

of an adolescent female (Burial 01-8) was documented (figure 5.8). She endured a puncture wound to the anterior surface of her first lumbar vertebra, consistent with a perimortem abdominal stab wound. Given the location and angle of entry, either or both the aorta and vena cava were perforated. Her body was not placed in this Muchik community’s cemetery but was carefully buried under the floor of a spacious, partially roofed ceramic firing area close to kilns (Shimada and Wagner 2007) and near offerings of fetal llamas. Despite their technical prowess, Middle Sicán craft production was immersed in mysticism (Shimada and Craig 2013). It is possible this young woman’s life was offered to ensure success in ceramic firing. Another inferred local Muchik group during the Middle Sicán era also conducted sacrifices. Rescue excavations documented a hilltop temple at Cerro Cerrillos (lower Reque drainage), consisting of a plaza and two

1 3 2   Klaus and Shimada

5.8. Huaca Sialupe Burial 01-8, an adolescent female buried in the floor of a spacious, partially roofed, ceramic firing area at the workshop. A perimortem puncture wound to the anterior surface of the L1 vertebra (arrow, right) is consistent with an abdominal stab wound. Drawing by Izumi Shimada and Cesar Samillán; photo by Haagen Klaus.

superimposed platforms (Centurión and Curo 2003). The smaller platform featured a roofed altar. Under the floor of the plaza and the large platform, the skeletal remains of at least 81 individuals were recovered. The human remains atop Cerro Cerrillos consisted of at least 53 subadults (between the ages of 2 and 18 years) and 28 adults (ranging from 20 to 35 years old) representing three or more depositional events (Klaus et al. 2010). Seventy-six percent of the primary burials and several

Bodies and Blood  1 3 3

of the disturbed contexts displayed perimortem mutilation. Cut marks produced by a metal blade on the anterior surfaces of cervical vertebrae indicate that the throats of at least half of the victims were deeply and repeatedly lacerated (figure 5.9a). In seven cases, mutilation of the throat was so extensive that partial decapitation ensued, and only the posterior muscles of the neck kept the head attached to the body. Cut marks were also observed on the anterior chest (clavicle, sternum, and ribs) in 22 of 32 of the most complete skeletons (figure 5.9b). These appear as initial incisions mutilating the anterior chest wall used to gain access to the thoracic cavity. These cut marks were associated with perimortem fractures of the posterior ribs in several individuals, indicating very forceful opening of the thorax. Despite such violence, the Cerro Cerrillos victims were carefully buried in accordance with traditional Muchik funerary rituals. Spondylus shells, Nectandra sp. seeds, and camelids accompanied a few victims. It is unclear if they had consumed Nectandra extract before death (known for its hallucinogenic, paralytic, and anticoagulant properties [Montoya A

b

5.9. Cut marks among the blood sacrifice victims of Cerro Cerrillos include (A) sharp force trauma, anterior cervical vertebrae, Burial 1; and (B) cut marks present on the anteriomedial surface of the left clavicle of Burial 18. Photos by Sam Scholes.

1 3 4   Klaus and Shimada

2004]). These victims were characterized by a high prevalence of enamel hypoplasias, anemia, and poor oral health—biological patterns that mirror their life history experiences and corollary health status with that of the commoner social strata (Klaus et al. 2010, in press). Victims’ dental phenotypic variation also mirrors other individuals of inferred local Muchik populations, such as at Illimo and Huaca Sialupe, and all appear to be local people. Late Middle Sicán Mass Human Sacrifice: Matrix 101 Sometime around AD 1050–1100, the Sicán political and religious leadership experienced an abrupt upheaval. The elite ancestor temples at Sicán and the auxiliary structures at their bases were systematically and intensely torched and abandoned. The Sicán Deity icon that permeated all artistic media nearly disappeared, and the capital was reestablished 12km to the southwest at Túcume (Shimada 2000). There is no evidence of foreign invasion, and no concurrent destruction was seen outside of the capital at commoner settlements. The Middle Sicán leadership appears to have met their end seemingly at the hands of its populace. Contributing factors likely included a multi-decade drought beginning around AD 1020, torrential rains, and the accompanying historic flood caused by a mega–El Niño event near the end of the drought (AD 1050– 1100). Effects of this mega–El Niño event have been documented over a large area of the Peruvian coast (Donnan 1990a; Moseley et al. 1983; Winsborough et al. 2012). At Sicán, the rains and flood left sediments up to 1m thick in and around the capital and changed the course of the La Leche River (Craig and Shimada 1986). Also, the aggrandizement of the Sicán elites and their ancestor cult was largely at the expense of local labor, well-being, and resources. This may have simultaneously eclipsed the tolerance of the people (Jennings 2009; Shimada 2000). The same El Niño event appears linked to a series of major sacrifice events recently documented at Sicán. In 2011, Museo Nacional Sicán archaeologists Carlos Elera, José Pinilla, and Ana Alva discovered a massive feature roughly at the center of the Great Plaza, equidistant between Huaca Las Ventanas and Huaca Loro. Designated Matrix 101 (figure 5.10), this feature was a multi-tiered, inverted cone more than 25m in diameter that tapered down to a central point 12m below the surface into the water table. Stratigraphy and artifact styles indicate this formation definitively dates to the Late Middle Sicán period

Bodies and Blood  1 3 5

5.10. The west sector of a section of Matrix 101, a mass burial of nearly 200 people located in the Grand Plaza between Huaca Las Ventanas and Huaca Loro. Associated with unprecedented rainfall during the Late Middle Sicán phase, our current working hypothesis is that it involved a ritual response to the mega El Niño that hit the coast between AD 1050–1100. Photo by Alex Bryce/Archives of Museo Nacional Sicán.

(AD 1050–1100) and appears to have been formed in three distinct but closely timed events. During the first event, a pit was excavated through a thick layer of recently deposited waterborne sediment, and at least 72 young and middle-aged adult males were placed around its deep central locus. Overlooking the north and east sides of the pit, an assemblage of large ceramic tinajas and porrones were placed into the ground. Such vessels were commonly used to brew or store chicha (maize beer) for large feasts. Later, Matrix 101 was reopened in a second event. They began by manipulating, removing, sorting, and reburying bones from most of the original burials. Experimental work by Toyne (2015a) shows that relatively advanced decomposition permitting such manipulation could occur in this environment in as little as 60 days. The disarticulated crania of at least 52 adult males were placed in a pile along with offerings of ceramic bowls, at least two dogs, and broken-off heads of several Sicán Deity bottles (figure 5.11a). Other deposits produced during the opening featured comingled collections of isolated sherds, broken bottles, camelid bones, and an estimated 1,000 ceramic heads depicting

A

b

1 3 6   Klaus and Shimada

c

5.11. Possible cases of object immolation. The head of a Sicán Deity vessel (A) was among hundreds of ceramics in Matrix 101 that seemed purposefully “decapitated” from their vessel bodies. This particular sherd was found in a cache of 52 human crania. Further examples of similar thinking may involve the bases of ceramic vessels (B, C) that were carefully punched out (in this case, a blackware vessel that accompanied Burial 86 at Huaca Las Ventanas). Photos by Haagen Klaus.

the Sicán Deity, humans, and animals, including llamas, felines, and monkeys, originating from fineware to utilitarian jars. These ceramic heads were so carefully broken off from their vessels that they appear purposefully “decapitated.” They join an increasing number of intentionally broken ceramic bottles where the center of their bases have been punched out in funerary contexts at Huaca Las Ventanas (figures 5.11b, 5.11c), evoking interpretations of killing or sacrificing a vessel. Once the previous skeletons were cleared out, a new group of at least 103 individuals was added into the deposit (again mostly adult males,

Bodies and Blood  1 37

but also including a few subadults and at least one adolescent female). Most bodies were positioned in ring-like clusters around the periphery of the center of the pit. Body positions included partially extended, flexed, hyperflexed, facedown, facedown flexed, facedown cross-legged, and haphazard (or “dumped in”) postures. The bodies were set on a hard, cracked, clay-like surface, signifying that exposed surfaces of Matrix 101 had become wet and then dried during the process of burial. It may have been raining at some point during the second event. Matrix 101 was then reburied. A third event transpired when sections of Matrix 101 were reopened, but this time only several sets of remains were altered; a few skulls were removed, while other skeletons were disarticulated by hand. Not long after, Sicán was set ablaze and abandoned. Overall, these events may have unfolded in as short as 6 to 10 months, well within the span of a mega–El Niño event, which can persist for two consecutive years. El Niño rainfalls are typically punctuated in intervals up to a few weeks, so it is possible to conduct the ritual activities discussed above during lulls, particularly if the previous performance was perceived as ineffectual in producing the desired effects (i.e., a return to climatic normalcy). Cut marks produced by throat-slitting were associated with only three individuals. The anterior surface of one isolated adult manubrium was affected by a probable perimortem puncture (stab) wound. A handful of others may have been bludgeoned to death, and one individual may have been impaled through the face and braincase (Hurtubise et al. 2014). Among the second group of bodies, telltale evidence points to the possibility of live entombment. Unusual hand positioning was seen in several burials, as in Burial 159, where the clasped left hand was drawn up to the face to cover the nose and mouth (figure 5.12a). In Burial 197, most of the articulated left hand was positioned inside the mouth. Burial 125 was a male in a facedown position, appearing to have fallen into his final resting place (figure 5.12b). The right forearm was positioned in front of the face, while the left arm was flexed out in front of the body. This is consistent with a reflex where the right arm attempts to shield the face from a face-first fall while the left arm attempts to brace the body. This individual probably perceived that he was falling but did not get up from where he hit the ground. In light of archaeothanatological perspectives (Duday 2009), these individuals appear to have been rapidly buried within a dense and hard matrix; such hand and limb positions appear to represent original positioning at the time of burial.

1 3 8   Klaus and Shimada

b

A

5.12. (A) Matrix 101 Burial 159, with the clasped left hand drawn up to the face as though to cover the nose and mouth (photo by Haagen Klaus). (B) Burial 125, appearing to have been thrown in their final resting place; positioning of the arms indicates a conscious attempt to brace from a fall. Photo by Alex Bryce/Archives of Museo Nacional Sicán.

Most individuals in Matrix 101 were not affected by episodes of childhood anemia and tooth enamel growth arrests (Klaus et al. 2013), which is the inverse of the pattern seen in contexts such as Cerro Cerrillos. The Matrix 101 individuals include the tallest adult males documented in these populations, suggesting they enjoyed at least adequate if not superior childhood nutrition. Their bones were mostly free of disease, and even minor expressions of osteoarthritis were rare. Oral health was overall excellent, indicating their diets did not contain large proportions of starchy carbohydrates. In essence, the people buried in Matrix 101 seem to have possessed multiple characteristics reflecting the lifestyles and diets of the elite social strata. Pending studies of DNA, dental phenotype, and biogeochemical variation will test the working hypothesis that the elite ethnic Sicán sacrificed more than 170 of their own during an unprecedented mega–El Niño event that preceded the downfall of the Middle Sicán polity and religion.

Discussion The long-term and cross-contextual study of burial samples from the Middle Sicán period in the Lambayeque valley shows that ritual killing

Bodies and Blood  1 3 9

transcended the political collapse of the Moche. Further, the frequency of sacrificial practices did not correspond to the notable decline of sacrificial iconography following the Moche. Instead, it seems ritual killing became more diverse. These observations lead to an emerging picture of a coexistence of various ontologies, as well as alternative constructions or experiences of personhood; sacrifice might not have been limited to a state monopoly. Human Sacrifice at the Sicán Precinct Ritual killing at the Sicán precinct surrounded three themes: dedication of sacred spaces, funerals of elite Sicán lords, and climatic disturbance. Middle Sicán dedicatory and retainer burials left no direct physical indications of violent death, and only a few individuals in the Matrix 101 mass killings were mutilated. Despite this fact, there is a very strong likelihood that these were all sacrificial deaths due to the simultaneous presence of several key lines of evidence. In a regional understanding of funerary patterns spanning the range of social variation, the placement of a body inside a corporate architectural construction, for example, is far removed from the cemeteries that received the vast majority of the dead. Age and sex variation (or lack thereof ) among retainers points to carefully targeted patterns of cultural selection (Coale and Demeny 1966). In terms of life history theory, these young people were in their biological prime, probabilistically the least likely persons to die a natural death. Blindfolds and cordage are nonexistent in cemeteries. As Middle Sicán funerary rituals were quite regimented, the departed were very carefully positioned. Facedown or otherwise haphazard body positions are equally absent in cemeteries. These observations converge on the likelihood that such burials involved unnatural death and distinctive meaning(s). Given the vagaries of soft tissue trauma, absence of evidence is not evidence of absence. Killing a retainer could have been conducted in a variety of ways that did not leave archaeologically perceptible traces. Also, the deaths of people destined to be offerings or tomb retainers probably were not intended as a blood sacrifice—but as an offering of an intact, relatively pristine body. A fatal dose of a toxin represents one possibility (Klaus 2009). Bodies placed within columnar sockets decomposed after the spaces had been filled with sand and gravel. Those people may have been buried alive. Functionally, blindfolds and bindings serve no need unless to hinder the

14 0  Klaus and Shimada

struggle of a notionally conscious individual. Interestingly, in Section A of the Denver Kero (Mackey and Pillsbury 2013: fig. 10), a seated, openeyed individual with outstretched arms embracing a cylindrical shape akin to a roof post) is depicted inside a small square chamber underneath the floor of a roofed temple-like structure. This could be a depiction of such a dedicatory burial in a subfloor columnar box. The two figures with outstretched arms beneath this individual may represent an act of receiving the sacrificial columnar box offering above. Overall, we are trying to avoid jumping to the conclusion that an unusual burial, or the presence of multiple bodies inside a single tomb, implicitly signals sacrifice. Instead, various lines of contextual evidence must be evaluated (Verano 1995). Similarly, employing contextual criteria of sacrifice, Eeckhout and Owens (2008) developed a model of a “potentially sacrificed individual” linking to the concept of “deviant” burial (Murphy 2008). Our approach is not dissimilar, but we wish to avoid the term deviant; we consider them within a spectrum of nonmodal or nonnormative practices. Perhaps their disposition was shaped most directly by immediate concerns or ideologically charged events—different from many other settings, but perfectly appropriate for the events at hand. Contextual clues point to a broadly conceived notion of Middle Sicán dedicatory and retainer burials as so-called invisible sacrifices that differed from Moche blood sacrifices. In future study, it is worth considering the possibility that the Sicán were expressing different concepts of personhood as compared to the earlier Moche. Were the Sicán emphasizing the sanctity of a “whole person” or the “whole victim” more so than the Moche, perhaps with the idea that only the whole individual (integrated body and soul) could serve the deities and ancestors? In the case of Matrix 101, “crisis sacrifices” in the Great Plaza may not have involved any need to dispatch the majority of victims in bloody rites, since there was already an excess of life-generating liquids (too much rain) and their deaths aimed to stabilize or reanimate some kind of cosmological center. However, the limited evidence of Moche dedicatory offerings (i.e., the children under Huaca de la Luna’s Plaza 3A [Bourget 2001a)] does suggest possible conceptual precedents. Sacrifice at Sicán in Diachronic Perspective Sacrifices at the Sicán precinct may demonstrate an increase of scale since Moche times. Dedicatory burials are known in a few Moche contexts. As just noted, construction of Plaza 3A at the Huaca de la Luna was

Bodies and Blood  14 1

immediately preceded by the offering of two children. At Late Moche Huaca Colorada (Jequetepeque valley), three women epitomizing the concept of a “foundation sacrifice” were thrown into the architectural fill (Swenson 2012). At the Late Moche urban settlement of Pampa Grande, a cache associated with the ramp leading to the Palatial Room Complex at Huaca Fortaleza contained the disarticulated bones of an incomplete child (Haas 1985). Additional subfloor columnar sockets atop the main body of Huaca Fortaleza mirror later Middle Sicán huacas. But at Pampa Grande, the only articles contained in the columnar boxes were llama offerings. By the Middle Sicán period, columnar box offering preferences shifted toward humans (paralleling Gaither et al., chapter 6). We can extrapolate a pattern based on the excavated subfloor columnar sockets atop various temple mounds at Sicán involving consistently alternating human and object offerings. At the ancestor veneration shrine atop Huaca Lercanlech, an estimated 250 people were buried among the 500 or more subfloor sockets. As noted, indications are that Huaca Loro, Huaca Las Ventanas, and the other major platform mounds at Sicán underwent parallel rituals (Shimada 1990; Shimada and Cavallaro 1986). Upward of 1,500 or more people may have been directly incorporated into six monumental huacas at Sicán. Considering the roughly contemporaneous construction of these shrines (ca. AD 1000), a profound expenditure of human life may have transpired over a relatively short period of time. Also, burials of young women or children within elite funerary contexts are well known from Moche settings such as Sipán or Úcupe (Alva and Donnan 1993; Bourget et al. 2012). Yet the greatest number of retainers ever found in a Moche tomb is eight. Incorporation of 23 such people in the Huaca Loro West Tomb is again consistent with an amplification of scale. The Middle Sicán people, as with the Moche, found themselves faced with the ecological, economic, and even cosmological disruption produced by a mega-ENSO event. In the AD 1050–1100 timeframe, another decades-long drought was followed by archaeologically documented flooding at the site of Sicán. Within this window of time, the Matrix 101 mass sacrifices transpired. In both Huaca de la Luna Plaza 3A and Matrix 101, sacrifice was associated with ENSO-related rain events. Moche victims at Huaca de la Luna were left unburied, and the bones were extensively manipulated during decomposition (Backo, chapter 3). At Matrix 101, bodies were very rapidly buried, and only later were remains manipulated. Sacrificial episodes at Huaca de la Luna Plaza 3A

14 2   Klaus and Shimada

sometimes involved as few as six people (Bourget 2001b). Matrix 101 received victims in two main episodes, involving deposition of 70 and 100-plus people, respectively. The El Niño–associated killings at Huaca de la Luna were sanguineous, whereas Matrix 101 involved mostly unmutilated bodies. Symbolic Associations at the Sicán Precinct Interpretation of the meanings of ritual killing at the Sicán precinct is challenging. Many past attempts implicitly conform to a concept of reciprocal obligation and social expectations. Such ideas may have indeed been a part of Andean thinking, but notions of “sacrifice as offering” seem too simplistic, envisioning the function or significance of sacrifice in a predetermined manner. Much value can be assigned to the death–fertility–regeneration concept (Arnold and Hastorf 2008; Mackey and Pillsbury 2013), but it may also walk a perilously thin line with lo andino (Andean uniformitarianism) presumptions. Such may gloss over the finer variability in human sacrifice, such as: meanings of why close kin were sacrificed in some high-status settings and not in others, the use of bodies versus body parts, as well as young-versus-old or femaleversus-male victims. Also, tomb retainers should not be relegated to a vague ancillary status, as they likely formed an integral, active part of the display and symbolisms carefully choreographed within each funerary context (see Brown [2010] and Kelly et al. [2008] for similar issues in Mississippian mortuary practices). There are likely a number of nested levels of significance that, when assessed and assembled, may build toward similarities with broader Andean ideas—but this is initiated from local and specific data, distinct from the top-down approach of lo andino analogies that may impose predetermined visions. This being said, human bodies in high-status tombs or sacred architecture can possess characteristics of “offerings.” Tomb retainers have long been assumed as functionaries, wives, and servants of the lord who followed him or her into the afterworld. Individuals who were carefully positioned nearby principal personages in elite tombs were often well dressed and wore ornaments, suggesting roles as companions or retainers who continued to serve the personage in afterlife. In addition, Ramírez (1998:224; also Shimada et al. 2004:384) relates the widely held notion of reciprocal obligations between the ruler or the ethnic leader (kuraka) and their subjects. The former cared for widows and retainers, who in turn were expected to accompany him in death and reciprocally fulfill obligations in the afterworld.

Bodies and Blood  14 3

However, some Middle Sicán retainers had been transformed into desiccated bodies by the time of interment. This highlights a process of planned collection and curation of bodies. We wonder if the Sicán held similar ideas about dead bodies as in other Andean traditions (i.e., the concept of mallqui and the mummified dead as a source of life-giving fertility). Were the corpses of Middle Sicán retainers thought to possess life-giving properties to sustain a lord in the afterworld or in the transformation into an ancestor? Mountains in many Andean cosmovisions are conceived of as animate entities that are the source of life-giving water (Burger 1992b). Similar ideas seem represented in both Moche and Sicán iconography, where the stepped pyramid icon (an encoded mountain) is depicted in murals, on painted bottles, sculpturally, and in metalwork alongside an ocean wave (encoded water), or with a wave emanating from the top of a stepped pyramid (Sharon 2001: fig. 3; Shimada 1995; Zevallos 1989). Moche and Middle Sicán adobe mounds were likely conceived of as artificial mountains in a kind of metonymy, drawing fundamental characteristics of mountains into the human realm where they could be manipulated (Bawden 1996:72). We think that some kind of mountain–huaca–water triad was represented directly by Huaca Loro. Support for this notion was found at the bottom of the north face of Huaca Loro, where a series of ceramic jars were carefully implanted at the mouths of narrow erosion gullies to collect rainwater cascading down their slopes (figure 5.13) (Matsumoto 2014; Shimada 2014a; Shimada and Elera 2007). A carefully constructed, narrow, sherd-lined canal found on the platform surface of the Huaca El Corte temple mound seems to have been used for pouring libations down the mound’s slope, akin to the movement of a river or rainwater (Shimada 1986). These cases would have communicated a visual symbolism of the huaca as a source of water, life, and fertility. Offering and burial of sacrifice victims in the architecture may have been linked to the life-giving forces of the huaca, perhaps sanctifying, stimulating, articulating, or sustaining its power. Human Sacrifice Outside of the Sicán Precinct The people of non-elite Middle Sicán social strata were no doubt more differentiated than we currently perceive, but the vast majority of them appear to have been affiliated with a broad-sense practice and identity of an ethnically Muchik substratum. As noted earlier, they continued a

14 4   Klaus and Shimada

5.13. At the bottom of the north face of the Huaca Loro temple mound, a series of ceramic jars were carefully placed at the mouths of narrow erosion gullies to collect rainwater cascading down the structure’s slopes. These appear to be part of a broader ritual, symbolic, and conceptual mountain–huaca–water triad represented directly by Huaca Loro. Photo by Izumi Shimada.

number of deeply rooted practices—including human sacrifice, albeit resituated within a new and distinct social reality. Evolution of Sacrifice from the Moche Culture to the Muchik Substratum Ritual killing associated with non-elite Muchik people during the Middle Sicán era carried forth various fundamental tenets of earlier Moche rituals, especially deep lacerations of the throat (Hamilton, chapter 2; Verano 2008a). Recent excavations at Huaca de la Luna Platform III show the innovation of a new element of ritual killing—mutilation and opening of the chest—which had its earliest known expression in the Late Moche era (Verano 2011). The common procedural threads of throat-slitting and chest-opening were evidently disseminated among the practitioners of ritual killing on the north coast via a process that has yet to be defined. These elements varyingly characterize basic trauma patterns at Huaca de la Luna, Dos Cabezas, Cerro Cerrillos, and settings even later in time (Klaus et al., chapter 7; Toyne 2011b; Toyne, chapter 8).

Bodies and Blood  14 5

If throat-slitting reflects a deep-seated regional ritual template, Muchik ritual at Cerro Cerrillos simultaneously involved ideas discontinuous with earlier examples. Moche rites often involved extremely violent acts (bludgeoning, decapitation, skinning, disarticulation, and genital mutilation) (Hamilton, chapter 2). At Cerro Cerrillos, mutilation was comparatively conservative. Individuals fitting a Moche victim selection profile (young and middle-aged men) were present at Cerro Cerrillos—but they were outnumbered almost two-to-one by children. Current data indicate that the Moche never violently mutilated the young. Also, there are no indications that these later victims were captives of any kind. Moche victims were sometimes dramatically dismembered (Backo, chapter 3; Verano 2008a) and deposited in mass graves. In contrast, victims at Cerro Cerrillos and Huaca Sialupe received burial treatments closely mirroring the funerals of almost any other commoner within the society apart from their location. They were wrapped in burial shrouds, sometimes provided food offerings (camelid body parts), and interred carefully in individual graves. Post-depositional offerings were even made to the young woman at Huaca Sialupe. She was among the remembered dead. This could represent a broader shift to a more generative kind of ritual violence (Duncan 2005). How much these differences reflect change over time remains to be seen. Still, late pre-Hispanic victim preference and treatment seem to represent a major discontinuity with antecedent rituals. Perhaps this also reveals regional differences in ritual unique to the northern north coast, though recent finds by Prieto et al. (2015) demonstrate that child victims, chest-opening, and careful burial were shared by Late Intermediate Period southern north coast sacrifices. Ideological and Political Dimensions of Muchik Blood Sacrifice Various sources describe blood and water as the fertile fluids necessary to sustain agricultural and mythic cycles within Andean traditions (Arnold and Hastorf 2008; Cieza de León 1553 [1998]). Cobo (1990 [1653]) indicated that human blood was widely perceived as a powerful (if not the most potent) substance to offer mountains and ancestors who engender fertility and water in the world. Application of bright-red cinnabar pigment to the face or body of a corpse in various Middle Sicán funerary contexts may have similarly symbolized blood to confer or communicate some kind of life force or vitality of the dead (Shimada 1995).

14 6   Klaus and Shimada

Just as sacrifices at the Sicán precinct appear linked to water and mountains, subaltern Muchik outside the capital seemed similarly concerned. Yet they articulated sacrifices differently. Unlike any of the adjacent hills and mountains, a naturally occurring spring is found a few dozen meters northwest of Cerro Cerrillos. Materialization of subterranean water may have strongly influenced the selection of this site for sacrifice. Also, a low cloud deck that produces mists or light rain sometimes enshrouds the summit of Cerro Cerrillos. It thus becomes tempting to consider how Cerro Cerrillos may have been perceived of in antiquity—especially considering accounts of mountains as conduits for water from a dark and wet underworld, where water percolates upward through mountains, forms into clouds, and produces life-giving rain (Arnold and Hastorf 2008; Arnold and Yapita 2000:344–353; Bastien 1978; Benson 2001; Classen 1993; Gose 1994; Sharon 2001). Children may have had a very special place in such conceptual systems. In eighteenth-century Mórrope, child sacrifice was a central element of a pre-Hispanic water myth (Rubiños y Andrade 1936 [1782]). Calancha’s (1972 [1638]) account of Chimú sacrifices of five-year-olds to the moon divinity Si was closely linked to water, blood, and the sea (Rowe 1948). Sources further afield and later in time depict children as “not quite human” in other Andean cosmologies (Sillar 1994), where infants originate as mountain spirits to occupy a liminal, fertility-inducing state. This allowed living children to be interlocutors on behalf of mallquis and huacas (Ramírez 2005:146). If even remotely similar ideas were present during the Middle Sicán era, the blood of children could have been a supercharged offering. Ritual killing also speaks to current thinking regarding the internal structure of Middle Sicán society. As noted, several lines of evidence indicate that some local communities in the Lambayeque countryside were semiautonomous Muchik collectives, such as Huaca Sialupe and Cerro Cerrillos. These locals did not simply “become” members of the new society or robotically participate in the religion of the new leaders. Not a single piece of material culture bearing the Sicán Deity icon was associated with either ritual killing or funerary activity at Cerro Cerrillos and Huaca Sialupe. The latter community was mass-producing—but not consuming—iconic and ideological emblems of the state. Ceramic assemblages at Cerro Cerrillos spanned local domestic wares, Mochicoid face-neck jars, and paleteada-impressed ceramics. There is also a

Bodies and Blood  147

shared lack of evidence of Middle Sicán administrative or supervisory presence at both sites. Perhaps the absence of apparent devotion or assimilation into the state religion reflected distinctions between the two groups vis-à-vis the politics of ethnicity. Perhaps there was a competitive dimension between Muchik and Sicán rituals, or perhaps differences in ritual practice reflected symbolic antagonism or ritualized resistance against a new order (as with burial rituals at Late Moche Galindo [Bawden 2001]). Alternatively, maybe the difference was pragmatic: elite Sicán ritual killing served the direct requirements of the Sicán ancestor cult, while Muchik ritual catered to the religious needs of local peoples. Whatever the state of things, not everyone appealed equally to the Sicán ancestor cult to renew life, water, and fertility in their world (Shimada 2014b). Sacrificial diversity may also be a side effect of elite Sicán strategies that pursued practical and integrative economic means to create a powerful political economy. Middle Sicán policies seem to have been more concerned with economic and social integration and tribute extraction rather than religious domination (Shimada 2014a; Shimada and Craig 2013). Strict ideological control or religious conversion of the local population, even in the Middle Sicán heartland, seems not to have been a goal of the elite. Diversity of sacrifice may reveal degrees of internal social and religious plurality (Klaus 2014b), just as the people at Huaca Sialupe seem to have enjoyed significant autonomy or exploited it themselves to pursue their own agendas (Shimada and Wagner 2007).

Conclusions and New Questions Late pre-Hispanic cultures such as the Middle Sicán lack a large corpus of naturalistic iconographic representations, making ritual killing less discernible and potentially hindering traditional approaches toward sacrifice. An effective countermeasure involves a long-term, cross-disciplinary investigation firmly embedded in regional, social, and historical contexts. The results of this approach revealed diverse forms of ritual violence from the capital precinct to local communities. Though there are socially based tendencies (many bloodless dedicatory offerings at the capital that contrast with mutilated victims elsewhere), we should still be cautious not to overly characterize these practices as solely elite- or commoner-associated traditions, as our own thinking might be typologically

14 8   Klaus and Shimada

constraining. It is important to remain vigilant and flexible regarding how human sacrifices may cut across social and ethnic boundaries and how their occurrence may relate to symbolic or social values attached to a given context or circumstance. As we continue to pursue the study of Middle Sicán sacrifice, we also remain wary of the assumption that ritual killing represented a simplistic form of human-supernatural reciprocity. Doing so could characterize sacrifice as the impulse of an uncritical, naïve, and superstitious populace trying to satisfy a one-dimensional need. We hope to outline a more encompassing framework that considers the ways in which Andeans attempted to articulate with, negotiate, and manipulate delicate and dynamic balances between life and death, the social and natural worlds, and their broader cosmos. Even though the sacrificial act itself could be short in duration and triggered by a single factor, it likely involved multiple social actors with diverse priorities, values, perspectives, and practices. Understanding these dimensions also entails elucidation of the social, natural, and historical contexts of a given sacrifice. These important themes are best explained through carefully coordinated and sustained interdisciplinary research and the careful use of analogy. Future excavations of Middle Sicán shaft tombs will enrich the current understanding of retainer burials. Possible loci of ritual killing include the tops of small mountains that rise from the valley floor where remnants of temple-like structures have been found. At Sicán itself, Huaca Botija holds significant interest. This artificially modified natural rocky outcrop is at the center of the sprawling Sicán precinct and may have been a huaca or a natural altar. As the study of Middle Sicán ideology, belief, and ritual progresses over time, our understandings of sacrifice will play a key role in deepening the archaeological scholarship on life, death, and history in ancient Peru. Acknowledgments Grants from the Shibusawa Ethnological Foundation, Tokyo Broadcasting System, the National Geographic Society, the Wenner-Gren Foundation, the H. John Heinz III Fund for Latin American Studies, and the Southern Illinois University Faculty Research Fund to I.S. are gratefully acknowledged for their support of the Sicán Archaeological Project from 1990–2014. The Unidad Ejecutora 005 Naymlap-Lambayeque funded the work at Matrix 101 (2011–2013), and generous assistance from Utah Valley University’s Office of Engaged Learning and College

Bodies and Blood  14 9 of Humanities and Social Sciences partially funded the Matrix 101 burial excavations and laboratory analysis. The 2003–2004 study of the human remains from Cerro Cerrillos was supported by The Ohio State University’s Department of Anthropology and Center for Latin American Studies. The authors are also indebted to the members of the Sicán Archaeological Project, Lambayeque Biohistory Project, and the Sicán National Museum. We are also grateful for the contributions, collaborations, generous data-sharing, and discussions over the years with Steve Bourget, Carlos Elera, Julie Farnum, Joanne Pillsbury, José Pinilla, Sarah Muno, J. Marla Toyne, John Verano, and Carlos Wester. J. Marla Toyne, Melissa Murphy, and three anonymous reviewers provided enlightening editorial critiques. Note 1. The term Sicán comes from the Muchik words Signan or Cican (House of the Moon or Temple of the Moon), local names for the ruins of the capital precinct (Shimada 2000). For a century or more the Sicán culture has been described by several inaccurate or problematic designations: Eten, Middle Chimú, and the Lambayeque culture, the latter referring to a stylistic characterization of looted and decontextualized funerary ceramics (Zevallos 1971).

1 5 0  Gaither et al.

chapter si x

Precious Gifts Mortuary Patterns and the Shift from Animal to Human Sacrifice at Santa Rita B in the Middle Chao Valley, Peru Catherine Gaither Metropolitan State University of Denver

Jonathan Kent Metropolitan State University of Denver

Jonathan Bethard Boston University School of Medicine

Victor Vasquez Arqueobios, Trujillo, Peru

Teresa Rosales Arqueobios, Trujillo, Peru

Archaeological studies of Andean mortuary patterns have been evolving in sophistication for many years, especially on the north coast of Peru—from the studies of elaborate high-status tombs (Sipán, Úcupe, Dos Cabezas, El Brujo) to community cemeteries (Pacatnamú, Huacas de Moche, and others) (Alva 1990; Alva and Donnan 1993; Bourget et al. 2012; Donnan 1973, 1986b, 1994, 1995, 1997b, 2001, 2003, 2007; Donnan and Mackey 1978; Millaire 2004; Nelson 1998; Verano 1995, 1997a, 2001a, 2001b). Additionally, notable examples of human sacrifice victims have been found at Huaca de la Luna, El Brujo, and Dos Cabezas (Bourget and Newman 1998; Cordy-Collins 2001a; Donnan and Cock 1997; Verano 2001a, 2001b, 2008a).

Precious Gifts  1 5 1

It is noteworthy that all of the above sites are major ceremonial centers, with artificially constructed huacas, elaborate painted murals, large populations, and other archaeological indicators of their occupants’ participation in an elite-driven form of what Donnan (2010) has called “Moche state religion.” Missing from the knowledge base regarding northern coastal mortuary practices, especially those associated with human sacrifice, are data from sites whose occupants were not living in centers but, rather, occupying urban settlements whose degree of incorporation within larger, dominant sociopolitical formations is archaeologically questionable. We refer to these sites as USUIs (Urban Settlements of Uncertain Degrees of Incorporation). Such sites (i.e., secondary- or tertiary-level settlements lacking monumental architecture) might be broadly analogous to some modern rural communities where religious centers carry on ritual activities on what would be considered sanctified ground but lack the grandiosity of the larger religious ceremonial centers. Consider, for example, the difference in ceremonies carried out at a rural Catholic church versus those at the Vatican. While the ideology behind the ceremonies is the same, there is a significant difference in the scale and elaboration of the ritual behavior. The ceremonies carried out at secondary or tertiary sites are not simply poor imitations of those rituals conducted at the larger ceremonial centers; rather, they reflect a foundational complex of ritual behavior that usually permeates every level of society—but that may be embodied, appropriated, or enacted differently on different levels of society. Thus, the ritual behavior associated with mortuary practices at a USUI might present a less grandiose version of what is seen at the larger ceremonial centers, thereby reflecting the consistency of the ideological foundation for the behavior. This model could also explain the modification of ritual as a reflection of the local-level concerns of the inhabitants of these types of sites. This might be expressed, for example, as a mixture of rites that includes the “state-level religion” and its associated activities, as well as local traditions, which may have persisted in the area through numerous political regimes. This makes study of USUIs crucial to the broader understanding of ideological paradigms that influence ritual activities. To adequately assess this issue, we must address the question regarding to what extent a Moche (or any) state religion permeated the belief systems of those living in USUIs. Dillehay’s commentary (2001:274) on the degree to which the Moche elite’s politics and religion penetrated

1 5 2   Gaither et al.

other areas and at different times is telling: “There is no doubt that there were long periods in Moche history when centralized authorities ruled from the city centers. My hunch is that during some periods of Moche history many commoners lived relatively free of state intervention, retaining some independence even during periods of strong centralized authority.” One of our goals in this chapter is to move the discussion beyond inference and provide relevant archaeological data on non-elite ideology and attendant ritual practices. Additionally, our data allow us to question the extent to which mortuary and sacrifice behaviors in post-Moche times (ca. AD 700–1532) represent a continuation of earlier patterns versus an introduction or adoption of foreign rituals or newly innovated local practices. Finally, the sacrifice of camelids and the role such behavior plays in ritual complexes and shifting ideological paradigms are topics that have received relatively little attention—especially in context with human sacrifice and mortuary treatment.

Background Recent work at the site of Santa Rita B on the north coast of Peru in the Chao valley has uncovered a sequence of animal and human sacrifices that may shed light on these issues.  Human sacrifice victims, apparently interred as offerings accompanying at least one, and possibly two, principal burials, overlie an archaeological layer that included at least eight intact camelid skeletons. The human skeletons were associated with a post-Moche culture that has been designated as a local Transitional Phase temporally equivalent to similar cultures of the northern Peruvian coast (Castillo et al. 2008) and possibly marked by socioeconomic decentralization and reorganization. Here, these skeletons date to approximately AD 1100 as determined by radiocarbon dates, whereas the underlying camelid skeletons appear to date to the Early Intermediate Period (ca. AD 200–750) as determined by associated ceramics and overlying, well-defined, Moche-period floors. The mortuary patterns observed in the Transitional Phase burials at Santa Rita B are consistent with the general pattern observed on the north coastal Peruvian region and in other areas as well. Specifically, this includes the sacrifice of non-adults and retainer sacrifices, in a pattern we have referred to elsewhere as “like with like,” and the ritual re-interment of individual body parts (Gaither et al. 2008, 2010). These general

Precious Gifts  1 5 3

patterns are seen in the Andean region spanning from at least the Early Intermediate Period, where examples of “like with like” can be seen at sites such as Dos Cabezas, where some of the so-called Moche giants may have been sacrificed because of their size similarity to the principal personage (Cordy-Collins and Merbs 2003; Donnan 2001), to the Late Horizon, where numerous sites have yielded both human and animal sacrifices associated with the Inka (Toyne 2011b). Notable among these are the capac hucha, or child sacrifices, found at high mountain sites (Chamberlain and Pearson 2001; Reinhard 1998; Vreeland 1998). Donnan and Foote (1978) describe an example of “like with like” from the Late Intermediate Period site of Huanchaco (ca. AD 1400). At this site, 17 child burials were uncovered, each of which was associated with at least one immature llama. The llamas were positioned directly over or beside the human bodies and exhibited flexed positions similar to all but four of the human burials. Retainer burials have also been demonstrated in numerous contexts, including the three tombs excavated at Huaca Rajada near the village of Sipán (Donnan 1995; Verano 1997b). Verano (1997b) discusses several apparent human sacrifices included with the principal burials. He describes the remains of at least three adolescents (15 to 18 years old at time of death) and one child (9–10 years old). Their seated or flexed burial positions, particularly that of the child (which was not a typical Moche burial position), indicate possible sacrifice. Numerous researchers have discussed the finds of what they refer to as secondary interments in funerary contexts in the Andean region. The finds range from the “jumbled bones” that Verano (1997a) refers to in the tombs at Sipán to ritual re-interment described by Millaire (2004). Nelson (1998) describes undisturbed burials containing incomplete skeletons at the Moche site of San José de Moro. Other researchers (Hecker and Hecker 1992; Klaus, 2003a; Klaus and Tam 2015; Millaire 2004) have described parts of bodies included in tombs with principal personages. Klaus and Tam (2015) and Klaus et al. (2013) note a preference in pre-Hispanic and Colonial funerary contexts alike for removal and reburial of long bones and skulls. For a more complete discussion on these Andean mortuary patterns, see Gaither et al. (2008, 2010). What is particularly important about the finds at Santa Rita B is that their occurrence at a USUI demonstrates ritual behavior. Although less grandiose, it is nonetheless consistent with rituals seen at larger ceremonial centers. One of the hallmarks of mainstream religions is the permeation of rituals throughout every level of a society. The occurrence

1 5 4   Gaither et al.

of these mortuary rituals at elite ceremonial centers as well as at USUIs such as Santa Rita B suggests that these practices are representative of a conservative and long-lived northern Peruvian coastal complex of ritual behavior. Additionally, the presence of camelid killings in the layers that temporally precede a sequence of human sacrifices may reflect a shift in either the intent of the sacrifice ritual as perceived by the practitioners, or in the social or political consequences of the ritual behaviors. This chapter will examine these patterns within the context of northern Peruvian coastal mortuary practices and within the broader context of sacrificial ritual in cross-cultural comparison. Furthermore, this work begins to address how these behaviors and changes in the ritual patterns might reflect shifting ideological or political paradigms.

Archaeological Context Santa Rita B is located roughly 86km south of Trujillo. The site sits in the lower part of the middle Chao valley at an average elevation of 420m above sea level. Although there is a central architectural area that probably served as an administrative core, there is little monumental architecture and no artificially constructed adobe platforms or truncated pyramid platforms, in contrast to civic-ceremonial centers such as Huacas de Moche, Sicán, Túcume, and Chan Chan. The site measures approximately 2.2km north-south and at least 2.0km east-west and is approximately 440 hectares in size ( 1/6 the size of the Chimú capital of Chan Chan). Most of the surface architecture is made of stacked stone or pirca construction, and its urban nature is clear, as there are more than 350 enclosures constructed using this technique, along with various water and debris flow control features (Brooks et al. 2001; Gaither et al. 2008, 2010; Rosales Tham 1999). With excavations completed in some parts of the site, the Santa Rita B Archaeological Project has been able to define the nature of the human occupation of the site and interpret aspects of its economic, social, political, and ideological history. Excavations over several years had focused on areas of apparent domestic architecture. One of these, known as Conjunto Arquitectónico 3 (or Architectural Complex No. 3, hereafter abbreviated as CA3) is a rockwalled compound measuring about 29m north-south by 25m east-west, subdivided into approximately 19 partly or completely enclosed spaces or rooms (see figure 6.1 for a map showing the location of CA3 within the site). In the upper strata of CA3, several human skeletons (figure 6.2)

Precious Gifts  1 5 5 Table 6.1. Calibrated Radiocarbon Dates from Human Bone Samples at Santa Rita B

Cat. No. Material

Sample

Type of Personage

Conventional Radio carbon Age (CRA)

Srb-560-1 Bone/R. Ribs Entierro #4 Principal Personage 900 + 40

Srb-561-1 Bone/L. Ribs Entierro #3

Probable Human Sacrifice

Probable Human Srb-562-1 Bone/R. Ribs Entierro #2 Sacrifice

890 + 40

850 + 40

C/12C

p†

13

2σ Cal*

–18.4

AD 1134–1271

0.886

AD 1046–1084

0.113

AD 1145–1272

0.928

AD 1049–1079

0.072

AD 1175–1281

0.976

AD 1162–1172

0.024

–15.5

–21.6

*2-sigma age ranges calibrated using the CALIB RADIOCARBON PROGRAM “SHcal 04” for the southern hemisphere developed by McCormac et al. (2004) used in conjunction with Stuiver and Reimer (1993). † Probability of actual calibrated date falling within stated range.

were encountered that produced calibrated 14C dates between AD 1050– 1289 (table 6.1). Underlying these deposits with the human remains was a layer of mixed debris containing Transitional Phase ceramics, Late Moche fineline sherds, various utilitarian plain wares, and disarticulated and butchered camelid bones. These were on top of a well-preserved floor, below which was a fill layer that included purely Moche ceramic debris, the most common being Moche IV sherds. Below that was another well-defined floor. It was below this floor that excavations revealed at least eight virtually complete and intact camelid skeletons that we infer were killed in a single event (see below). One additional complete camelid skeleton was found just east of the eastern wall of the CA3 enclosure. Figure 6.3 illustrates the camelid skeleton locations.

Osteological Analysis (Human Remains): Methods Sex and Age Estimation Sex estimation was accomplished only for individuals whose skeletal maturation permitted the visual assessment of sexually dimorphic pelvic or cranial characteristics (i.e., Buikstra and Ubelaker 1994; Phenice 1969; White et al. 2012). Supplementary long bone measurements

1 5 6   Gaither et al.

6.1. Map of the location of CA3 within the site of Santa Rita B. Field topographic data and Autocad map by Topografo Cesar Valverde S.

included femoral and humeral head diameters (Bass 1995) and can be more accurate for sex estimation than cranial morphology, particularly in small-bodied populations such as Andeans (Spradley and Jantz 2011). Age estimation was accomplished in adults utilizing well-documented age-related changes of the auricular surface of the ilium, the sternal rib extremities, the pubic symphyseal face, and cranial suture closure (Bass 2005; Buikstra and Ubelaker 1994; White et al. 2012). These features were assigned a score or phase with an associated age

Precious Gifts  1 5 7

range and mean age. To avoid the potential artificial narrowing or expansion of possible age ranges produced by some multifactorial age estimation techniques (see Digangi and Moore 2013; Uhl and Nawrocki 2010), the simple average technique was utilized here. It is calculated by adding the mean ages of each of the scoring techniques and dividing by the number of techniques used. This technique tested relatively well in Uhl and Nawrocki’s (2010) study, correctly aging 55% of individuals; it simply produces an average numerical value. Uhl and Nawrocki (2010) suggest using a ± 10-year range to accompany this value, but depending on what characteristics are used for age estimation, error ranges may be unacceptably large in many cases. Therefore, once mean ages were calculated from these three multifactorial techniques, those means were averaged and a more standard ± 5-year age range was utilized if several skeletal features were available for age estimation. If skeletal elements were missing and prevented the utilization of one or more of the techniques above, the age range was increased to ± 10 years. Age estimation in non-adults was accomplished utilizing tooth crown and root formation, tooth eruption, long bone lengths, and epiphyseal

6.2. Map of the location of the human skeletons within CA3. Field map by J. Kent, skeletal placement by R. Busch.

1 5 8   Gaither et al.

6.3. Map of the location of camelid skeletons within CA3. Field map by J. Kent, camelid placement by K. Bedingfield.

union (Buikstra and Ubelaker 1994; Gaither 2007, 2004; Scheuer and Black 2000; Ubelaker 1999; White et al. 2012). Designation of an individual as non-adult is based on the union of the spheno-occipital synchondrosis (or basilar suture) and third molar formation and eruption. Published age ranges, which are typically smaller for non-adults than those established in adults, are given as appropriate and are dependent upon the technique used. Analysis of Skeletal Pathology and Trauma Traumatic and pathological lesions were analyzed in accordance with established methods (Buikstra and Ubelaker 1994; Lovell 2008; White et al. 2012). The specific location of the lesion, including the bone, side, and aspect, was meticulously recorded, and the timing of the injuries was conservatively determined after careful consideration of the osseous evidence, burial context, and taphonomic processes (Berryman and

Precious Gifts  1 5 9

Haun 1996; Berryman and Symes 1998; Buikstra and Ubelaker 1994; Sauer 1998; Sledzik 1998; White et al. 2012). More specifically, antemortem abnormalities were diagnosed on the basis of macroscopic evidence of healing. This includes porosities near the lesions that indicate bone activity and resorption, rounding of the edges of a lesion, or the presence of a callus or bone remodeling (Sauer 1998). The determination of perimortem fractures was based on the presence of characteristics that indicate “fresh” (i.e., moisture-laden) bone and by taking into consideration known taphonomic processes on bone (Crist et al. 1997). Specifically, lesions exhibiting one of the following were characterized as perimortem in origin: (1) radiating and/or concentric fracture(s); (2) hinge fracture(s); (3) greenstick fracture(s); (4) sharp edges that do not demonstrate evidence of remodeling; and (5) an angled shape often with a surface that may also demonstrate delamination (beveling or peeling) and/or knapping or flaking (Sauer 1998). Additionally, in all cases, the color of a fracture margin had to be consistent with the color of the rest of the bone surface. Fracture lines that were lighter in color indicate postmortem breakage (Buikstra and Ubelaker 1994). It is important here to note the effects of taphonomic processes on non-adult bone and their relevance with respect to distinguishing perimortem trauma from postmortem breakage. This study took into consideration the structural characteristics of non-adult bone versus adult bone, the evidence for the rate of loading forces, and the criteria associated with fresh bone in the attempt to diagnose perimortem trauma as distinct from postmortem breakage. The behavior of bone when loading forces are applied depends on age and architecture in conjunction with the rate of the force applied. The latter is due to the viscoelastic nature of bone. The viscous element of bone allows it to yield to applied force and stretch prior to fracturing when the rate of the force is slow. This results in a ductile response demonstrated by plastic deformation. When the loading is faster and more acute, the viscous element does not have time to yield to the force, and the result is a brittle response involving pathological overload, fracture, and production and displacement of numerous fragments (Crist et al. 1997; Pierce et al. 2004; Rockwood and Green 1984). Age-related differences in structural characteristics of bone play a role in this process as well. It has been determined that Haversian canals are wider in children, thereby occupying a greater portion of the cortex. This factor makes young bone more porous and more flexible than adult bone, allowing for absorption of relatively (though not absolutely) more

1 6 0  Gaither et al.

energy before either permanent deformation or fracture results (Currey and Butler 1975; Pierce et al. 2004; Rang and Wenger 2006). Pores in the cortex limit the extension of fractures in the same way that a hole drilled through the end of a crack in a window will prevent the crack from extending (Rang and Wenger 2006). The lower mineral content of young bone as compared to adult bone, however, means that it is less dense and therefore weaker in general than adult bone (Pierce et al. 2004). Additionally, it can fail under both compressive and tensile loading forces, whereas adult bone tends to fail only with tension (Crist et al. 1997; Pierce et al. 2004; Porter 1989). Though stronger overall, adult bone is less elastic than non-adult bone and typically exhibits a more brittle response. However, non-adults also demonstrate a brittle response with higher loading rates (Crist et al. 1997; Currey and Butler 1975; Rang and Wenger 2006). Specifically, with respect to cranial bones, Crist et al. (1997) concluded that antemortem (including perimortem) fractures would likely cross fracture lines and continue onto the adjacent bone, whereas postmortem breakage resulting from compressive forces generated by burial would likely result in radiating fractures that terminate and dissipate along sutural articulations. This is due to the fact that the cross-sectional area of cranial bones is very small and therefore the stiffness of the bones is reduced. Their study also noted that children are less likely to demonstrate displaced fractures because their periosteum is stronger than that of adults and therefore less easily torn. Research has demonstrated that the periosteal sleeve is much thicker in children than adults, and this acts as a restraint to fracture displacement (Currey and Butler 1975; Rang and Wenger 2006). Given this information, it is easy to see that trauma patterns in non-adult bone may demonstrate significantly different characteristics than adult bone. Understanding the biomechanical stressors and osteological responses in both non-adult and adult bone is important for the interpretation of the nature and timing of trauma identified in bone, and it was a guiding factor in our diagnostic interpretation of perimortem trauma in the remains at Santa Rita B.

Results of Osteological Analyses Inferred Principal Burials CA-3 has thus far yielded the remains of at least eight individuals. Five are believed to be sacrifices, one is consistent with a case of re-interred

Precious Gifts  1 6 1

body parts, and two are possible principal burials. The individuals described in this section were determined to be possible principal, or primary, interments based on the following criteria: intentional cranial modification in the form of symmetrical parieto-occipital flattening, supine extended burial positions (with arms extended alongside the body in the case of Burial 4) without any indication of haphazard interment, and inclusion of associated camelid sacrifices aligned with or near the human body and oriented in a similar pattern (i.e., directional head/ feet orientation). It is also consistent with body positions observed at San Jose de Moro during the Transitional Phase (Rucabado and Castillo 2003; Tello et al. 2003). Here, Burial 4 (figure 6.4) was a child found in a prepared grave in the form of a shallow trench dug into the alluvial matrix from which the body was recovered. (The other skeletons were simply on top of the alluvial layer.) While no nonorganic grave goods associated specifically with these interments were recovered, the camelid and human sacrifices could be considered grave offerings. This complete camelid, which was less than 3 months of age, was immediately to the right side of the child, with the right arm of the child underneath the body of the animal. The camelid was also oriented on the same north-south axis in a supine position. Additionally, trauma could not be identified on these remains, although there were cranial lesions consistent with chronic childhood anemia, a common finding in Peruvian coastal populations (Blom et al. 2005). While they could conceivably have been killed utilizing a method that would not leave visible marks on the skeleton, it would represent a significant divergence from the consistent pattern seen with other individuals designated as sacrifices. Based on all of this, Burial 4 is likely a principal interment. Burial 8 (figure 6.5) is identified only as a possible principal interment. Burial 8 was the partial skeleton of an adult male, aged 32 ± 5 years at death. The partial cranium of an adult (SRB-04-18, FS 50) recovered from a disturbed context also demonstrates intentional cranial modification in the form of symmetrical parieto-occipital flattening. This cranium may be associated with this individual. The two were in close proximity to one another, and the looter’s pit that disturbed the upper portion of the body of Burial 8 contained the modified cranium. The partial cranium was of an adult, and the only other nearby burial was that of a non-adult. The body of Burial 8 was buried in an extended supine position, suggesting differential treatment when compared to

1 6 2   Gaither et al.

6.4. Inferred principal personage (Burial 4). Photo by Catherine Gaither.

Precious Gifts  1 6 3

6.5. Possible second principal personage (Burial 8). Photo by Catherine Gaither.

the majority of those individuals demonstrating clear evidence of perimortem trauma and who were buried usually prone or flexed. Thus, this context may represent a second principal burial, but it is not possible to definitively identify it as such. The remains of a camelid were also found next to this skeleton and in the same position as this individual (i.e., in orientation of the head and feet; also see figure 6.9 below). Human Sacrifices and Re-interred Body Parts The sacrifice victims include individuals who demonstrate injuries indicative of perimortem trauma. These include Burial 1, a probable female aged 15 years ± 36 months at death who sustained possible perimortem injuries to the cranium. This included two areas of perimortem blunt force trauma as evidenced by well-defined points of impact (one unhealed injury on the left parietal bone, and another on the left side of the frontal bone), linear radiating fractures, which are frequently associated with intracranial hematomas and, in this case, crossed suture lines (Galloway 1999), adherent fragments of bone (i.e., hinge fractures), and consistent coloration of the fracture lines as compared to the

1 6 4   Gaither et al.

6.6. Several traumatic injuries on human skeletons of Santa Rita B: parry fracture, Burial 2 (top left); cut mark on sternum, Burial 2 (bottom left); cut marks on ribs and a severed rib end, Burial 5 (top right); and perimortem femoral fracture, Burial 9 (bottom right). Photos by Catherine Gaither.

surrounding bone. Burial 2, an adult male aged 30 ± 5 years at death, demonstrated perimortem cut marks on the sternum and near the sternal end of the sixth left rib. Designation of this feature as a perimortem injury was based on a lack of healing and consistent fracture margin coloration. Additionally, this individual demonstrated a perimortem parry fracture of the right ulna. Parry fractures are often considered defensive wounds and are a likely interpretation in this case given additional evidence of trauma associated with interpersonal violence (Lahren and Berryman 1984; Lovell 1997; Walker 2001). Perimortem designation was based on the oblique nature of the fracture in combination with delamination (also known as beveling or peeling) of the fracture edges. Burial 5, a non-adult aged 10–12 years at death, demonstrated perimortem cut marks on the sternal ends of the left sixth and seventh ribs. The sternal end of the sixth rib was completely severed. Again, a lack of healing and consistent coloration of the fracture margins indicated these injuries occurred in the perimortem interval. Burial 9, a non-adult possible male aged 12 years ± 30 months at death, also demonstrated

Precious Gifts  1 6 5

possible perimortem injuries to the cranium, including a fracture to the occipital bone with resulting fragmentation of the bone and the production of a linear radiating fracture onto the right parietal bone. This individual also demonstrated a perimortem fracture of the right femur. The fracture was oblique with delaminated edges demonstrating consistent fracture line coloration, and the surrounding bone demonstrated a discoloration that could be consistent with staining from hematomas (Sauer 1984). Figure 6.6 details several of the skeletal traumatic injuries. These injuries suggest significant trauma in the perimortem interval, and the parry fracture found on Burial 2 suggests the possibility that while significant physical violence surrounded their deaths various individuals may have fought back and resisted when possible. Aside from the perimortem injuries, a non-funerary treatment of the dead consistent with sacrificial ritual was also inferred from haphazard atypical body positions. For instance, the skeleton of Burial 2 was in a supine flexed position with one hand in front of the body and the other behind his back. Burial 3 was in a prone position with the hands drawn up under the body in a manner consistent with breaking a fall. Burial 5 was in a supine position with the arms extended above the head as indicated by the position of the humeral heads, and Burial 9 was in a prone position with large rocks placed on top of the body (figure 6.7). These body positions are not consistent with any identified funerary practice from this, or any, time period in the Andean region. Body positions such as that seen with Burial 3 appear to indicate a sudden, violent death, particularly given that the position is consistent with breaking a fall. It is also evident that the individual did not move from that position after falling (compare with Klaus and Shimada, chapter 5, figure 5.12b). In addition to haphazard body positions, there are articulated remains of disembodied body parts that do not demonstrate any evidence of cultural or natural transformation processes that might have disturbed the remains. These include the remains from Burial 10 (figure 6.8) and one partially articulated left foot not associated with a numbered burial. While we cannot prove these represent sacrificed individuals given the lack of traumatic evidence, their inclusion in a funerary context with the remains of individuals who do demonstrate evidence of sacrifice suggests the possibility they may have suffered a similar fate. Additionally, these remains are suggestive of ritual re-interment of body parts, which is another long-term north coast pattern present at this and other sites (Millaire 2004; Shimada et al. 2015).

1 6 6   Gaither et al.

6.7. Burial positions of sacrifice victims: Burial 9 (top left); Burial 3 (top right); Burial 5 (bottom left); and Burial 2 (bottom right). Photo by Catherine Gaither.

6.8. Burial 10, an undisturbed articulated partial body. Photo by Catherine Gaither.

Precious Gifts  1 67

The inclusion of sacrifice victims in proximity to individuals interpreted as principal burials suggests they were intended as retainer sacrifices. Additionally, the bodies demonstrate articulated undisturbed skeletons, and there were no blowfly puparia or other insect remains found in association with any of the bodies, suggesting they were buried very shortly after death. Insects have been demonstrated to interact with human corpses in a predictable manner (Byrd and Castner 2001), and on the north coast of Peru the environment preserves insect exoskeletal remains when present, owing both to the arid climate and the composition of chitin, a resilient long chain polymer derived from glucose (Gilby and McKellar 1970). Thus, body positions, traumatic injury patterns, and evidence of rapid postmortem interment are consistent with an interpretation of burials of principal individuals accompanied by sacrifice victims in a funerary context rather than other possible explanations, such as a massacre or some kind of natural catastrophic event. Isotopic Chemistry Isotopic analyses conducted on the human remains at Santa Rita B suggest that the sacrificed individuals came from within the community (Bethard et al. 2008). We were able to obtain usable collagen from five individuals (Burials 2, 4, 5, 8, and 10) for carbon and nitrogen isoto15 13 pic analysis. The results demonstrate relatively low δ N and high δ C levels consistent with C4 maize consumers utilizing terrestrial protein resources. Strontium analysis performed on Burials 1, 2, 3, and 9 is also consistent with the local signature for the Chao valley as determined by levels present in local modern fauna. This presents strong evidence that the sacrificed individuals were local residents consuming a similar diet as those individuals identified as principal interments. The Camelid Sacrifices Beneath the stratum that included the human bodies, excavations revealed mixed debris, including Transitional Period ceramics, sherds of Late Moche ceramic vessels decorated in characteristic fineline painting, various utilitarian plain wares, and disarticulated and butchered camelid bones. These remains were lying on top of a well-preserved floor that extended across the complex and in some areas beyond its currently standing pirca stone walls. Below this floor was a fill with purely diagnostic Late Moche ceramic debris, which was underlain by

1 6 8   Gaither et al.

6.9. One of the Early Intermediate Period camelid sacrifices at Santa Rita B.

a second floor level. It was underneath this lower floor that excavations revealed complete camelids. Figure 6.9 illustrates an example of one of the camelid skeletons. These animal skeletons were virtually intact. Several of them exhibited perimortem blunt force trauma to the head. One had a cut mark on the sternal portion of the fifth rib. At least eight intact animals were exposed during excavation. Most were in front of the large elevated platform labeled in figure 6.3 (i.e., between the entrance to the large room in which the platform is situated and the platform itself ). In every case, the camelids were oriented with the head to the east, facing in the direction of the highlands. Some were lying on their left side, while others were in the camelid rest position with the forelimbs tucked in backward under the chest cavity. Given these patterns of trauma and the careful arrangement of the bodies, this set of remains may confidently be interpreted as animal sacrifices. Additionally, the consistent orientation and stratigraphic position of the animals suggests this represents a single event and definitively preceded the laying down of the lower, Late Moche–period floor.

Precious Gifts  1 6 9

Discussion and Interpretation Most Andeanists see some degree of cultural continuity among inhabitants of different time periods on the north coast of Peru.  The Transitional Period mortuary patterns seen at Santa Rita B are consistent in certain ways with patterns seen at other north coast sites, including ceremonial sites that are contemporaneous, earlier, and later in time. A comparison of broader funerary rituals with the finds from CA3 reveals four common patterns that persist from the Early Intermediate Period through the Late Horizon. These include: (1) retainer sacrifice; (2) child or non-adult sacrifice; (3) a principle we refer to as “like with like,” which involves a metaphorical equivalence between the sacrifice and the individual with whom the sacrifice victim is included; and (4) the ritual re-interment of body parts (see Gaither et al. [2008] and Gaither et al. [2010] for a more complete discussion of these patterns through space and time, as well as their relevance to the interpretation of the remains at Santa Rita B). Table 6.2 lists the sites, time periods, and associated mortuary patterns. Despite the clear diversity in the expression of some patterns, these finds support a notion of at least some level of cultural continuity. In fact, we argue that much of the diversity represents variations of the broader themes enumerated in this chapter. Additionally, the finds here demonstrate that mortuary behavior seen at large ceremonial centers was also taking place at USUIs. This suggests a widespread, deeply embedded, and ritually reinforced ideology—one that resulted in the permeation of ritual behavior to every social strata and one in which the basic ritual requirements persisted through time. For example, the different scale of the activity at Santa Rita B is demonstrated by the lack of designated monumental ritual space. The sacrifices took place in an area of domestic architecture rather than a huaca or other clearly delimited sanctified ground. For the community at Santa Rita B, CA3 served a dual social purpose, much as how a school gymnasium might occasionally be utilized as a meeting hall. At larger ceremonial centers, there are specifically designated areas for ritual activities that are constantly maintained as such, but USUIs might not be able to spare the space and/or labor for such a luxury. Therefore, the community utilized various spaces for multiple purposes, both sacred and secular. The finds at Santa Rita B do, however, prompt the question as to why there would be a major shift

17 0  Gaither et al. Table 6.2. Documented Mortuary Patterns at Andean Sites Through Time* Cultural Period/Site

Mortuary Patterns Child Sacrifice Retainer Sacrifice “Like with Like” Ritual Re-interment

Colonial (AD 1533– ) San Pedro de Mórrope

X

Late Horizon (AD 1450–1532) Mt. Nevada de Ampato X Machu Picchu X Ethnohistoric data X X X Middle Horizon/Late Intermediate Period (AD 800–1400) Huanchaco X X Huaca Cao Viejo X X Huaca Loro/Huaca las Ventanas X X Lengache X X Santa Rita B X X X Early Horizon/Early Intermediate Period (900 BC–AD 200) Dos Cabezas X X Sipán X X Huacas de Moche X X X San José de Moro Maranga Complex X Paracas X X X

X

X

X X X

* Compiled based on: de Herrera 1944; Donnan and Foote 1978; Moseley 1992; McEwan and Van de Guchte 1992; Kent and Kowta 1994; Franco et al. 1994; Donnan 1995; Reinhard 1996; Reinhard 1997; Verano 1997; Bourget and Newman 1998; Reinhard 1998; Vreeland 1998; Bourget 2001; Donnan 2001; Chamberlain and Pearson 2001; Verano 2001a; Cordy-Collins and Merbs 2003; Donnan 2003; Kent et al 2003; Verano 2003; Shimada et al. 2004; Kent et al. 2006; Gaither et al. 2008; Klaus 2008; Gaither et al. 2010; Koschmieder and Gaither 2010.

from animal to human sacrifice given the almost certain biological and cultural continuity between the people of the late Early Intermediate and Transitional Periods. A broader examination of relevant non-Andean data on human and animal sacrifices and the reasons behind these behaviors reveals numerous interesting possibilities. Some of these might be applicable if one recalls that at USUIs (by definition, given that their degree of incorporation is unknown) there is no necessity for all the canons of any state

Precious Gifts  171

religion to be followed. This would especially be true, for example, if the cultural identity of the occupants of Santa Rita B was made up at least partially of herders. Indeed, there are overwhelming data (Kent et al. 2003; 2006) suggesting camelid agropastoralism was carried out during the Early Intermediate Period and in subsequent phases of the occupation of Santa Rita B. This differs from Moche ceremonial sites that depended more on marine subsistence and irrigation agriculture (see such a characterization of life in Moche ceremonial centers by Quilter and Koons [2012], who do not incorporate considerations of pastoralism in their discussion of the Moche polity). Bulliet (2005), for example, discusses human and animal sacrifice in the context of transitions from hunter/foragers to the “predomesticity” of herders. Basing his argument on both ethnographic and archaeological data, he postulates that as herding becomes more central to various aspects of life animal sacrifice is seen as being presided over by a sort of big man (“big shot”) who “orders the sacrifices performed on behalf of his people . . . and [allows them to] share in the distribution of meat from the sacrificed animal” (Bulliet 2005:122). It has not been ascertained whether the Santa Rita B camelid or human sacrifices were presided over by such a person. Certainly, the cultures involved were likely sufficiently hierarchical to have someone or a small group of people function in this role. However, these cultures were also certainly more complex than the “big man” or big woman level of social organization, thereby placing sacrifice well within the sphere of domesticity rather than predomesticity as posited by Bulliet (2005). Additionally, as far as the cut mark and traumatic evidence suggests, neither humans nor camelids were butchered for meat distribution. Although there is some evidence of postmortem or post-depositional dismemberment of some of the humans, this appears more consistent with north coast mortuary rituals where graves were revisited and reopened and the remains manipulated (see Gaither et al. 2008; 2010). It is most doubtful, therefore, that the objective of sacrifice at Santa Rita B was feasting that entailed acquisition of human or camelid meat through ritual killing. Ritual killing of these people associated with the funerals of the people interred in Burials 4 and 8 remains the most consistent explanation. There is also the question of why there was a shift from animal to human sacrifice. We argue that the shift does not indicate cultural discontinuity given the maintenance of the underlying patterns, the fact that in both cases something precious is being offered, and the fact that

17 2   Gaither et al.

the same area was being utilized for both animal and human rituals. There are numerous examples in religious history where acceptable substitutions are made for ritual purposes without a significant change in ideology. For example, such a tale from Christianity and Islam involves the substitution of a sheep in lieu of the sacrifice of Abraham’s son. All that is necessary is that the substitution be acceptable to the cosmological force(s) receiving the offering. In this Andean case, since the change involved moving from animal to human sacrifices, the substitution may have even been argued as mandated by gods or ancestors. The answer may lie in the intent of the sacrifice as perceived by the practitioners or in the social and political consequences of the sacrifices; of course, both possibilities should be examined. The discussion of the intended functions of sacrifice in anthropology has a venerable history. Bourdillon and Fortes (1980) summarize several ways for approaching sacrifice in a cross-cultural vein useful to anthropologists as well as theologians. Only a few need be mentioned here; they include: Calculated sacrifices: This is conceived of as propitiatory gift-giving to deities, ancestors, or other beings and forces, perhaps as a way to communicate with the supernatural or even to create an obligatory reciprocity between the living and the forces in their cosmos. The logic of this kind of ritual involves giving up and destroying or killing valuable objects, animals, or people in hopes that supernatural forces will favor the world of the living in return. Firth (1963a) has also discussed the related concept that sacrifices may be seen as an offering by people of resources that are valuable in hopes of obtaining goodwill from the gods. This could include, in this Peruvian case, balance in terms of lifegiving rainfalls and protection from either too much rain (El Niño events) or too little (La Niña events). Interestingly, much like the mass killing that occurred at the distant Sicán capital (Klaus and Shimada, chapter 5), the human sacrifices at Santa Rita B occurred following a 50-year mega– El Niño event that began around AD 1050, and we wonder if this factor played a role in the evolution of ritual in this community. If the camelids of Santa Rita B were seen as precious during the Late Moche period (and work at Moche V Pampa Grande in the Lambayeque valley indicates camelids were central to the economy and foodways [Shimada 1994a]), the symbolic cost of giving up multiple precious animals as gifts may have been perceived as being balanced by a gain in some kind

Precious Gifts  173

of supernatural favor. Similarly, Transitional period occupants of Santa Rita B may have believed that offering the lives of a few young individuals would ensure communal well-being after the death of an individual of high status. Prestige sacrifices: A second way of explaining these ritual deaths is that sacrifice can serve as a source of prestige for those conducting the rituals (Bourdillon and Fortes 1980). These are designed to enhance the status of the person, persons, or groups responsible for them, usually because it may be viewed as an act of generosity. Such a possibility could explain the Late Moche period camelid sacrifices, if they were ordered by an individual or group seeking enhanced social standing; the Transitional Period human victims may be a reflection of the need to demonstrate the high status of the deceased principal personage or, alternatively, could have served the purpose of solidifying the prestige of a newly appointed political or religious leader as part of a transition in power upon the death of the previous political or religious leader. Klaus (2013, personal communication) suggested that this interpretation may be particularly apropos, considering that the Transitional Period was an era of sociopolitical decentralization and a generalized “power vacuum” on the southern north coast of Peru. In a time when many different local lords and foreign polities such as the Wari, Cajamarca, and Pachacamac cultures may have been vying for influence, violent sacrifice accompanying the funerals of important personages may have been able to demonstrate local authority in a public display.

Although this is certainly possible, it is unclear whether the agropastoral occupants at Santa Rita B were governed by local lords vying for power with other, yet to be specified groups. Alternatively, expression of influence could have been focused for immediate local consumption within Santa Rita B. Sociopolitical cohesion: The killings at Santa Rita B may also be intended to promote cohesion through sacrifice (Bourdillon 1980). If the community as a whole participated in the sacrifice, the shared ritual would serve to unite the sacrificers. This is seen in other pastoral cultures as described by Fortes (1980). This would be especially

174   Gaither et al.

desirable in times of social stress. The expansion of the Moche into the middle Chao valley in Middle Moche times may have occasioned this sort of social strain, and the act of sacrifice could bring locals into the new order. This would be useful if the Moche administrators were attempting to bring pastoralists into a semi-hegemonic polity. In a similar manner, the human sacrifices of the Transitional period may have been a way for the new leaders to establish and unite the community. The isotopic analyses offer support for the hypothesis that the sacrificed individuals came from within the community (Bethard et al. 2008). This may reflect participation in the ritual behavior, and support for the new regime, by the community as a whole, even to the point of giving up young lives. Though separated by both thousands of miles and a thousand years, Late Horizon capacocha sacrifices could reflect similar thinking (de Herrera 1944; McEwan and Van de Guchte 1992; Reinhard 1996, 1997, 1998). With this general understanding of some likely reasons behind ritual behavior involving sacrifice, a look at the social and political consequences of sacrificial acts is possible to achieve archaeologically. In CA3, there was a shift from Late Moche practices involving sacrificed camelids to post-Moche (or Transitional Period) rituals involving sacrificed humans. How does this compare with information on the Late Moche– Transitional Period mortuary behavior elsewhere in the north coast region? And under what conditions might a similar shift have occurred? Fortunately, comparative perspectives emerging from the Jequetepeque valley provide some clues. In attempting to elucidate changing ritual practices at Jequetepeque regional sites, particularly at San José de Moro, Castillo et al. (2008) provide details on numerous cemeteries and tombs from both Late Moche and Transitional occupational phases. An Early Transitional tomb (Tomb M-U615), dated to after the Late Moche collapse (post–AD 850), is described as consisting of successive funeral events, with bodies deposited in superimposed strata with fill in between and accompanied by various offerings. Fifty-eight interred individuals in four major phases of deposition were found. Each phase in turn was made up of several episodes of burial and re-deposition of bodies. Indeed, such a communal tomb differs starkly from earlier Moche practices on the north coast. In fact, the authors note that there seems to be a deliberate attempt by Transitional Period people to reject some aspects of Moche funerary cannons even while preserving the complexity of previous Moche

Precious Gifts  175

political organization (Castillo et al. 2008; also Tomasto et al., chapter 11). If people at Santa Rita B were participating in a similar rejection and reformulation, then the shift in sacrifice may well reflect that. In other words, the shift away from camelids toward humans as sacrifice victims may constitute evidence of a change in the southern reaches of the Moche region reflecting different symbolisms for the same message or, conversely, different symbolisms embodying a new message. Either way, this can be a difficult change to make. Paradigm shifts, particularly those involving deeply held, long-term religious beliefs, can be culturally destabilizing if they happen too rapidly or without enough support, much along the lines of what Bawden (2001) saw as contributing factors in the final collapse of Moche V Galindo. Preserving the basic ideological foundations, albeit in an altered state, may have been a way to achieve the desired result. Rejection of the earlier practices allowed for the establishment of a new order without wholesale rejection of the religious infrastructure that served to validate the power of the incoming regime. In fact, a shift to human sacrifice as indicated at Santa Rita B could have served to solidify the power of the new system. If the leaders attempted to propagate a message that they were stronger or bore greater legitimacy than their discredited predecessors, they offered something even more precious—human lives as opposed to camelids. General yet instructive analogies along these lines can be derived from examples elsewhere in the global archaeological record. Voigt (2012) argues that the inclusion of sacrificed women and children in the Galatian levels of the archaeological site Yassıhöyük (generally accepted as historic Gordion) in Turkey represents attempts by a conquering regime to subjugate the local population. She also notes that European texts link Celtic human and animal sacrifices to periods of social stress. Hesse et al. (2012) interpreted donkey sacrifices at Tel Haror in Israel as evidence of the value of donkeys as more than mere beasts of burden, analogous to the llama and alpaca in the Andean world. Additionally, Hesse et al. argue that the sacrifice of “precious gifts,” such as valuable animals and even humans, represent “scripts” embedded with deeper cultural meaning in relation to the sacred. They argue that differences in sacrificial rituals indicate a shift in the cultural script, one embedded in repetitive actions—a structuralist interpretation that would suggest that it represented core cultural values. Thus, in the case of Santa Rita B, the shift from animal to human sacrifice suggests a significant change in the superficial cultural script, the

17 6   Gaither et al.

impetus for which might be the legitimization of a new regime. More to the point, Weber (2012) argues that the mortuary rituals apparent at Tell Umm el-Marra in Syria, which included “death pits” containing hundreds of human and animal retainer sacrifices, served to “reconstitute and re-legitimize the ruling regime” (Weber 2012:182). According to Jing and Flad (2005), during the Shang Dynasty (ca. 1600–1046 BC) in China, the animals used in sacrificial activities changed over time. They note that pigs, dogs, and cattle were sacrificial victims in the late Neolithic period, whereas horses and sheep were added later in the Shang period. The reason for the shift was a change in the way elites controlled their world and enhanced their power. “Animal sacrifice in particular was an important aspect of the process by which elite power was constructed” (Jing and Flad 2005:252). Such a legitimization of power can be facilitated by manufacturing strong ties between the living and the dead, something that might be accomplished through sacrifice—particularly if the community willingly provided victims. Carter (2012) argues that a lack of self-interest on the part of the donors honors the spirits and/or gods associated with their earthly representatives. If we speculate somewhat and apply similar ideas to Santa Rita B, then we can postulate that the community leaders during Late Moche times may have been doing the same with camelids—that is, using them as a means of exerting authority or enhancing or legitimizing power. By the Transitional Period, the killing of humans replaced the sacrifice of animals. This may have represented a new kind of control, one that was intent on demonstrating and legitimizing elite power through deeply held and widely shared forms of ritual.

Conclusion The mortuary patterns at Santa Rita B are consistent with general long-term ritual patterns in this region of the world, reflecting a deeply embedded and persistent ideological paradigm that included child (or non-adult) sacrifice, retainer sacrifice, “like with like,” and ritual reinterment of body parts. The layers of camelid sacrifice are not inconsistent with that worldview. Rather, they represent something precious that was offered as part of that ideology, perhaps also serving the purpose of enhancing the power of those in control. At the same time, these rituals seem to have been situationally dynamic. The shift from animal to

Precious Gifts  177

human sacrifice may represent a shift in power, one that maintained the long-held ideological infrastructure even as it superficially rejected the perceived cultural traditions of a previous political/religious system and worked toward establishing a new and more stable order. Acknowledgments We would like to sincerely acknowledge the following people and institutions for their invaluable assistance: Haagen Klaus, J. Marla Toyne, three outside anonymous reviewers, Carol Mackey, John Verano, Manuel Fernandez, Jeffrey Quilter, Orlando Archibeque, Jorge Chiguala A., Santiago Uceda C., the 2002, 2004, 2005, 2006, and 2007 field personnel, Stacy Greenwood, Aiden Kent, the California Institute for Peruvian Studies, the Universidad Nacional de Trujillo– La Libertad, and Metropolitan State University of Denver (MSU Denver). MSU Denver also provided partial funding for the project that was financed in large part by participating field personnel.

17 8   Klaus et al.

chapter seven

Human Sacrifice at the Chotuna-Chornancap Archaeological Complex Traditions and Transformations of Ritual Violence Under Chimú and Inka Rule Haagen D. Klaus Department of Sociology and Anthropology, George Mason University

Bethany L. Turner Department of Anthropology, Georgia State University

Fausto Saldaña Museo Nacional de Arqueología y Etnografía Hans Heinrich Brüning de Lambayeque, Peru

Samuel Castillo Museo Nacional de Arqueología y Etnografía Hans Heinrich Brüning de Lambayeque, Peru

Carlos Wester Museo Nacional de Arqueología y Etnografía Hans Heinrich Brüning de Lambayeque, Peru

Beginning in the latter part of the Late Intermediate Period through the Late Horizon (AD 1350–1532), ancient Peru saw the rise and fall of states and empires. During this era, the north coast was transformed in a multitude of ways. The strategic Lambayeque Valley Complex came to be ruled by the foreign Chimú and Inka empires. It was a time of unparalleled change. Paradoxically, twentieth-century archaeology paid mostly cursory archaeological attention to the Late Intermediate period north coast. This led to persistent but questionable impressions of regional conflict and foreign domination—and it was long assumed that this was a time and place lacking the social, ideological, and artistic dynamism of preceding eras (Shimada 2000: 110). To establish a better understanding of late pre-Hispanic Lambayeque,

Human Sacrifice at Chotuna-Chornancap  17 9

the Museo Nacional de Arqueología y Etnografía Hans Heinrich Brüning initiated a long-term study of the Chotuna-Chornancap Archaeological Complex in the lower Lambayeque valley. During the first four years of the project, two settings of human sacrifice were discovered. Here, we present a multidisciplinary and comparative bioarchaeological study of these contexts. This chapter examines mortuary patterns, biological stress, trauma, and isotopic variation to discern the identity of the victims, ritual structure, and the meanings and motivations of these acts. These killings also shed light on the broader development and variation of sacrifice in the late pre-Hispanic era while highlighting interplays between local religion and styles of imperial rule.

The Late Pre-Hispanic Backdrop The Chimú Conquest of the Lambayeque Valley As the Late Sicán/Lambayeque polity emerged circa AD 1100, the Chimú kingdom also began to rise (Moseley and Day 1982; Moseley and Cordy-Collins 1990). This predatory imperial state was centered at the Moche valley site of Chan Chan. Covering some 25 square kilometers, this city was the nexus of a vast redistributive tribute and trading economy. The Lambayeque Valley Complex was conquered during the third wave of Chimú expansion circa AD 1375 (Mackey 2011). Containing roughly a third of the population and a third of the arable land on the north coast, Lambayeque was a bonanza to the imperial treasury and more than doubled the empire’s productive resources. Chimú administrators pushed the extent of cultivable land in Lambayeque to its preHispanic maximum. Administration was accomplished via a three-tiered administrative hierarchy (Shimada 2000: 103). Túcume was co-opted as the provincial capital as immense amounts of labor and resources were invested into the site (Heyerdahl et al. 1995). Additional Chimú footprints are seen in the form of abundant Chimú ceramics, free-standing rectangular masonry compounds, and U-shaped audencias at locations crucial to resource control (Fernández 2004; Shimada 1990). Potentially coercive Chimú power is hinted at by construction of substantial hilltop fortified compounds at Jotoro, the Pampa de Chaparrí, and along the Taymi canal intake where control over irrigation water was exerted (Hayashida 2006; Tschuaner 2001). The Chimú focused on politics and economy rather than remaking local culture or religion. Regional burial patterns reflected continuation of earlier, ethnically Muchik rituals at La Caleta de San José (Rodríguez

1 8 0  Klaus et al.

1995), Jotoro (Martínez 2011), La Pava (Fernández 2011), Úcupe (Wester 1996), and Huaca Sontillo (Klaus 2003b). Local paleteada pottery production continued, and Tschauner and colleagues’ (1994) excavation of a specialized ceramic workshop at Pampa de Burros 10km east of Cinto revealed the production of Chimú utilitarian ceramics without on-site supervision. Autonomous residents also carried out distribution of their wares outside of the redistributive Chimú economy. The Inka Conquest Around AD 1470, the Chimú engaged in direct conflict with the Inka and were definitively vanquished (Cabello Balboa 1951 [1586]: 319). One of the most significant impacts following incorporation into the Inka province of Chinchasuyu involved land tenure (Ramírez 1990: 519–525). Land was set aside for centrally controlled state production and access to hunting grounds, fishing waters, forests, and mines; some agricultural communities were reorganized in the strategic Pampa de Chaparrí (Hayashida 2006). The Inka desired control over the unique coastal system of economic exchange (Ramírez 1990) and found value in Chimú administrative approaches. If local paramount lords were compliant, then their offices were respected, but any resistance was forcefully countered (Cabello Balboa (1951 [1586]). Túcume remained the paramount provincial center, and the site thrived. Buried in Huaca Larga (Room 1, Platform 2), a funerary bundle including accouterments of Inka nobility was suggested to have been “the last Inca governor of Túcume who controlled the entire Lambayeque region” (Narváez 1995b: 96). Local funerary and sacrificial traditions persisted at the Temple of the Sacred Stone (Toyne 2011b; Toyne, chapter 8). Hayashida’s (1998) work at the Tambo Real and La Viña workshops found them to be associated with administrative centers in the La Leche drainage. Potters were clearly producing polychrome wares imitating Cuzco motifs blended with traditional north coast conventions. The Chotuna-Chornancap Archaeological Complex During this era of foreign occupation, another monumental site held great sway: the Chotuna-Chornancap archaeological complex (figure 7.1), located among the monte scrub vegetation and sand dunes of the lower Lambayeque river drainage. Chotuna has long been associated

7.1. The principal monumental constructions of the expansive Chotuna-Chornancap Archaeological Complex, consisting of the Chotuna and Chornancap sectors. Huaca Chornancap is 1km to the west of the Chotuna sector and Huaca de los Sacrificios. Map by Roy Gutierrez, Museo Brüning.

Human Sacrifice at Chotuna-Chornancap  18 1

1 8 2   Klaus et al.

with the Naymlap legend (Cabello de Balboa 1951 [1586]; Rubiños y Andrade 1936 [1782]). This oral tale tells of the reign of a cultural hero at the site of Chot (inferred as Chotuna). Some consider the legend as authentic (Fernández 2012; but see alternative positions by Rowe [1946]; Shimada [1990]; and Zuidema [1990]). Despite tantalizing archaeological explorations of this topic (Donnan 2012), future work can design testable hypotheses to better evaluate its veracity. Though Chotuna and Chornancap were known since the late nineteenth century (Brüning 1922; Kosok 1965), the first large-scale excavations were conducted by Christopher Donnan from 1980–1982 (Donnan 1989, 1990a, 1990b; 2012). In 2006, Brüning Museum archaeologists initiated the current project (Wester 2010). Site chronology is based on multiple radiometric dates (Donnan 1990b; Trimborn 1979) and temporally discrete artifact stylistic variation. Aside from some Late Moche ceramic sherds, the earliest structures at Chotuna-Chornancap appear to be Middle Sicán/Classic Lambayeque (AD 900–1100). The Chimú occupation (ca. AD 1375–1470) involved a massive influx of material and human capital that underwrote a major expansion to the point that Chotuna-Chornancap was likely the principal secondary satellite to Túcume. At Chornancap, nobility were interred within the sacred precinct (Wester 2012). Locals continued to bury their dead at Chotuna even into the Early Colonial period (Donnan 2012). The Chotuna sector is 8km from the city of Lambayeque and 4.5km from the Pacific shoreline. It has been extensively covered by windblown Aeolian deposits and covers some 20 hectares. It is dominated by six monumental adobe brick platform mounds, including Huaca Chotuna and Huaca Susy, and are surrounded by multiple smaller temples decorated with murals, palaces, plazas, walled enclosures, craft workshops, cemeteries, domestic occupations, and ancillary facilities (Donnan 2012; Wester 2010). The Chornancap sector is 1.5km east of Chotuna. Huaca Chornancap is a three-tiered, truncated adobe platform 70m long, 50m wide, and 25m high, forming a t-shaped configuration. It was surrounded to the north and south by expansive low platforms and walled plazas containing temples, murals, and royal courts (Wester et al. 2010b; Wester 2012). This sector burgeoned during Late Sicán/Lambayeque and Chimú occupations, but by the time of the Inka, Chornancap was abandoned. Within a 5km radius of the center of the complex, there are at least 100 medium- and small-scale sites with Middle and Late Moche, Middle

Human Sacrifice at Chotuna-Chornancap  18 3

Sicán, Chimú, and Chimú/Inka occupations. Chotuna-Chornancap was the center of gravity of the lower Lambayeque drainage for several centuries. While we are just beginning to clarify the ethnicities and identities of the people at Chotuna-Chornancap and its environs, an overwhelming Gallinazocoid and Muchik stylistic presence is noted in the ceramic assemblages (Donnan 2012: 125–165).

Materials and Methods Among the mortuary contexts documented at the Chotuna-Chornancap complex, two settings of human sacrifice were uncovered from 2008– 2010. The first context featured at least 24 sacrificed individuals placed into Huaca Chornancap’s North Platform during the Chimú occupation. The second group of at least 33 Inka-era victims was placed in one of the smaller mounds of the Chotuna sector: Huaca de los Sacrificios. In diachronic, cross-contextual, and bioarchaeological frameworks, we explore four questions related to these contexts. First, what do the victims lives before death reveal about the sacrificial process? Second, why were these people selected to die? Third, what evidence exists regarding their manner(s) of death? And fourth, what did their burial process entail? We characterize the victims’ lives in terms of the skeletal “imprints” of lived experiences (Gowland and Knüsel 2006; Sofaer 2006) through patterns of skeletal biology and morbidity (due to space limitations, see Buikstra and Ubelaker [1994], Klaus and Tam [2009b, 2010], and Klaus et al. [2009] for descriptive protocols and differential diagnoses) (figure 7.2). Demographic variation is described in terms of age (based on dental development and long bone epiphyseal fusion in subadults and the morphology of the pubic symphysis, auricular surface, and cranial suture fusion in adults) and sex (based on standard morphological variation of the adult os coxae and skull). Morbidity at different points of life history was gauged by patterning of: (a) enamel hypoplasias, produced in episodes of childhood metabolic stress that results in bands of decreased enamel thickness (Goodman and Rose 1991); (b) porotic hyperostosis lesions of the cranial vault indicative of chronic hemolytic and megaloblastic childhood anemias (Walker et al. 2009); and (c) infectious disease, ranging from specific skeletal syndromes (tuberculosis, treponemal disease) to chronic infections by nonspecified pathogens that produce osteomyelitis and periostitis. Inferences about physical activity were derived from the prevalence

1 8 4   AKlaus et al.

d

b

c

7.2. Some of the key skeletal variables used to assess morbidity and life history among the sacrifice victims at Chotuna-Chornancap: (A) enamel hypoplasias, Huaca de los Sacrificios Burial 9; (B) healed porotic hyperostosis lesions of the cranial vault, Huaca de los Sacrificios Burial 20; (C) dental caries, Chonancap Norte Burial 10; and (D) abnormal new bone formation on the surface of the tibiae, Chornancap Norte Burial 4. Photos by Sam Scholes.

of degenerative joint disease (DJD) (Klaus et al. 2009). Victim’s diets were reconstructed through the study of linked pathological phenomena of dental caries, antemortem tooth loss (AMTL), and calculus formation (Hillson 2008), the prevalence of which are tethered to the long-term contributions of starchy carbohydrates to a diet (Larsen 2015). Biogeochemical variation was examined to longitudinally reconstruct dietary composition and residential mobility (Turner et al. 2013). First, we isolated rib bone carbonate and collagen from 29 individuals at Huaca de

Human Sacrifice at Chotuna-Chornancap  18 5

los Sacrificios to examine carbon (δ13C) and nitrogen (δ15N) isotope variation. Due to the nature of lamellar bone remodeling, these data reflect average diet over the last decade or so of life. Second, these values were compared to δ13C and δ15N values from keratin in sufficiently preserved hair from 15 individuals to characterize their diet in the last weeks to months of life. Third, oxygen (δ18O) isotope values of rib bone are linked to the geographical variation of the isotopic composition of imbibed meteoric water and can help assess an individual’s geographical origin and mobility in the years prior to death. For a detailed description of the chemistry and analytical methods used in the isotopic analysis of these remains, see Turner et al. (2013). All skeletal elements were scored for evidence of trauma in terms of the location, type, shape/size characteristics, and degree of healing. These data characterize ante- and perimortem injuries as well as patterns of perimortem mutilation. The living disposed of sacrificed bodies in ritualized burial. Descriptions of mortuary patterns included observations of body and limb positioning, use of grave goods, and evidence of post-burial rites. A taphonomic analysis rooted in an anatomical or archaeothanatological approach allowed the reconstruction of mortuary programs in light of decompositional processes and bone positioning (Duday 2009).

Results Chimú-Era Sacrifice: Chornancap Norte (~AD 1375–1470) On the central-west side of the Chornancap Norte north platform (figures 7.3, 7.4) were 18 individual burial pits placed into the architecture during the Chimú occupation. Demography, Skeletal Biology, and Health Status Among the 18 pits, skeletal remains of at least 24 individuals were documented. Most skeletons were well preserved and complete. Ten individuals were subadults under 21 years of age. The remainder included six adult males and three adult females. Using the summary age statistic (Lovejoy et al. 1985; Klaus 2008a), further demographic breakdown of the Chornancap Norte sample was established using standardized age class distributions (Klaus 2008a). One child was between 0 and 4 years of age, four subadults were between 5 and 14 years of age, six adolescents/young adults were between 15 and 24 years of age, four

1 8 6   Klaus et al.

7.3. The North Platform of Huaca Chornancap, with the T-shaped platform mound of Huaca Chornancap in the background. The North Platform featured plazas, polychrome murals, an inferred altar or shrine, and an audencia-like space with an apparent throne. The Chimú era sacrifices were placed in the platform’s southwest quadrant (arrows). Photo courtesy of Carlos Wester, Museo Brüning.

adults were between 25–34 years, and two adults were between 35 and 44 years of age. All victims at Chornancap Norte were skeletally healthy at the time of death, but many showed a history of morbidity earlier in life. Enamel hypoplasias were observed in 8/15 (or 53%) of individuals with preserved dentitions. Porotic hyperostosis lesions were observed in 10/15 (or 66%) of the victims. Periosteal infection was present in two skeletons. Burial 8 possessed inactive (healed) periosteal lesions on the plural surfaces of multiple ribs consistent with chronic respiratory infection. Despite the older ages of the adults, only two adult males suffered from minor DJD in their thoracic and lumbar vertebrae, knee, and foot joints. The subadults were characterized by excellent oral health, but they were young, so oral pathological conditions had little time to progress. The children skew overall crude prevalence of oral disease downward for the entire sample. Nonetheless, patterns of oral health are similar to other late pre-Hispanic Lambayeque regional populations (Klaus and Tam 2010), with overall frequency of dental caries present in 7/141 (or

Human Sacrifice at Chotuna-Chornancap  187

5%) of observed anterior teeth (incisors and canines) and 45/194 (or 23.1%) of observed posterior teeth (premolars and molars). Antemortem tooth loss was uncommon in the anterior dentition (3/141 tooth loci, or 2.13%) but more common in the posterior dentition (26/195 tooth loci, or 13.4%). One abscess was observed, and 60% of the individuals had minor to moderate accumulations of dental calculus. A

d

c

b

e

7.4. Sacrificed individuals interred at Chornancap Norte; pictured here are Burial 2 (A), Burial 7 (B), Burial 14 (C), Burial 15 (D), and Burial 19 (E). The victims were positioned in variations of extended and seated/flexed postures. Grave goods were rare, but note the half-shell of a Spondylus sp. bivalve placed with Burial 19. Photos by Fausto Saldaña and Samuel Castillo, Museo Brüning.

1 8 8   Klaus et al.

Traumatic Injury Sixteen of the Chornancap Norte skeletons were complete enough to evaluate trauma (figure 7.5). Nine individuals at Chornancap Norte had their throats slit (figure 7.6a). Seven individuals displayed evidence of chest mutilation (figure 7.6b, figure 7.6c). Throat and chest lacerations co-occurred in five individuals. The narrow V-shaped cross-sections of all cuts are consistent with a metal blade. Cuts were documented in various combinations on the anterior bodies of the cervical C1, C2, and C3 vertebrae. Trauma in three cases included bilateral cuts into zygopophyseal joint spaces (the anterior surface of the inferior articular facet) of C2 vertebrae consistent with incomplete decapitation (figure 7.7). In the thoracic regions, cuts were observed on the right clavicle, manubrium, right ribs, and one T8 vertebra. One manubrium was bisected on its left lateral side (figure 7.6c). Chornancap Norte Burial 16 was decapitated. The head was placed by the left side of the seated body with the C1–3 vertebrae still attached to the basicranium. The anterior surface of the T8 vertebra of Burial 10

7.5. Composite map of the cut marks documented in all the individuals recovered from Huaca Chornancap’s North Platform. Image by Roy Gutierrez and Haagen Klaus.

A

Human Sacrifice at Chotuna-Chornancap  18 9 7.6. Examples of sharp force trauma among the Chornancap North Platform burials: (A) cut marks on the anterior surface of the C2 vertebra of Burial 16; (B) multiple fine cut marks on the anterior surface of the manubrium of Burial 7; and (C) a left lateral view of the same manubrium, showing that it had been bisected. Photos by Sam Scholes.

b

c

7.7. Cut marks in the zygopophyseal joint spaces of the C2 vertebra, Chornancap Norte Burial 1. Photo by Sam Scholes.

1 9 0  Klaus et al.

featured distinct horizontal cut marks. These marks originated deep within the thoracic cavity potentially as the byproduct of excision of internal organs. A pair of cut marks on the posteriomedial aspect of the first right rib of Burial 10 also suggest a cutting motion originating from within the thoracic cavity. Mortuary Patterns and Burial Taphonomy Of the 18 burials, three had been disturbed by later activities (see below). Of the remaining contexts, eight contained bodies in an extended position, sometimes with the legs and arms slightly flexed. In Burial 14, the positioning of the hands is consistent with being bound at the wrists. Five other individuals were placed in a seated/flexed position, and an additional two bodies were hyperflexed on their backs with their legs splayed open. All but one victim was positioned to face west, facing the nearby sea. Grave goods were rarely provisioned at Chornancap Norte. Three burials included only Spondylus shells, and Burial 13 clasped a shell in their hands. Camelid elements (mostly toes) accompanied two other burials, and in one context a stirrup-spout vessel was placed with a victim. There were no perceptible correlations between age, sex, burial position, or grave goods. Anatomical positioning of long bones, downward collapse of the thoracic cage, and intact finger and toe phalanges point to burial early in the postmortem period (Duday 2009). Given that preservation conditions were quite good, the absence of insect remains (Huchet and Greenberg 2010) independently demonstrates rapid interment. Yet the living came to revisit these burials. Remains within three burial pits had been incompletely exhumed in antiquity. Additional bones and bone fragments of at least six individuals were found under several primary burials, indicating these grave spaces were recycled or reutilized for later interments. The Inka Era Sacrifices at Huaca de los Sacrificios (ca. AD 1470–1532) Some 240m to the north of Huaca Chotuna is the comparatively inconspicuous adobe mound of Huaca de Los Sacrificios (previously designated Huaca Norte; Wester et al. 2010a). It measures approximately 26m north-south, 24m east-west, and 28m high. Its construction spans four phases beginning in the Middle Sicán (or Classic Lambayeque) period and concluding with the Inka occupation. The entire structure seems exclusively linked to ritual activities. The remains of 33 individuals in

Human Sacrifice at Chotuna-Chornancap  191

7.8. The sacrificed individuals in Chamber 1, Huaca de los Sacrificios, with the inset illustrating the location of this triple burial atop the structure. Photo by Haagen Klaus. Bottom right: a small Inka-style conopa copper figurine placed by the bodies. Photo by Fausto Saldaña, Museo Brüning.

1 9 2   Klaus et al.

7.9. A downward-facing view of Chamber 2 in the process of excavation. It had been opened in antiquity, and the remains of four of the five individuals in the shaft-like enclosure were altered and manipulated. Photo by Fausto Saldaña, Museo Brüning.

three adobe brick chambers were found within the final architectural phase. Chamber 1, measuring 2m x 3m, contained the remains of two adolescent women and a child (Burials 1–3) (figure 7.8). A small, male conopa copper figurine of diagnostic Inka design was placed by the bodies along with seven subadult camelids. The adjacent Chamber 2 was a shaft-like enclosure 2.3m x 1.8m and nearly 3m deep (figure 7.9). Chamber 2 contained the remains of five individuals (Burials 4–8) corresponding to two subadults, two young adult males, and one middle-aged adult female. Chamber 2’s stratigraphy reveals at least four distinct depositions and

Human Sacrifice at Chotuna-Chornancap  193

reopenings, leading to removal of the skulls of Burials 4, 5, and 8, and the right hand and both feet of Burial 8. Also, the legs of Burial 6 were eventually replaced but haphazardly re-interred. Chamber 3 was a multilevel, multiple burial context (figure 7.10). The first level contained 10 individuals (Burials 9–18) who were subadults 4 to 15 years old. Eight more skeletons (Burials 19–26) were found directly beneath them. Given the intimate stratigraphic association and lack of disturbance of the deeper burials, it is likely all 18 bodies relate to a single depositional event. Some 50cm below Chamber 3, two adobe compartments were found, likely representing a slightly earlier episode during the Inka phase (figure 7.11). In the north compartment, two adult women and a child (Burials 27–29) were propped up against the east wall. In the adjacent south

A

7.10. Chamber 3 at Huaca de los Sacrificios was a multilevel grave. The first level (A) contained 10 subadults between 4–15 years old. Directly beneath them (B) eight more skeletons were encountered. Photos by Fausto Saldaña, Museo Brüning.

b

1 9 4   Klaus et al. 7.11. Chamber 3 North Compartment (A) and South Compartment (B) containing the bodies of children and women. Photos by Fausto Saldaña, Museo Brüning.

A

b

Human Sacrifice at Chotuna-Chornancap  195

compartment, remains of one adult woman and three children (Burials 30–33) were crammed into another small enclosure. Both groups faced west. Demography, Skeletal Biology, Health Status, and Isotopic Data Of the 33 individuals buried at Huaca de Los Sacrificios, the assemblage included 22 subadults between 4 and 14 years of age (66%), four adolescents between 15 to 20 years (18%), and seven adults ranging from 21 to 34 (31.8%). Of those old enough to estimate sex, seven were probable or definite females, and three were probable or definite males. The Huaca de los Sacrificios victims showed no evidence of active disease states, but they experienced morbidity earlier in life. Enamel hypoplasias were present on antimeric anterior teeth in 24/32 (or 73%) of the individuals. Porotic hyperostosis lesions were observed in 13/30 (or 43%) of the 30 observable cranial vaults. Abnormal periosteal new bone formation was present in the tibiae of three individuals. Three skeletons demonstrated ante- and perimortem traumatic injuries. Burials 3 and 22 had broken ribs. In the case of the former, the incomplete degree of healing indicated the injury to the posterior left fifth rib was sustained probably a month or two before death. The unhealed perimortem fracture of the left radius and ulna of Burial 33 could represent a possible defensive wound (parry fracture) incurred at or around the time of death. Minor expressions of DJD were observed in five of the seven adults, present only in the shoulder and knee joints. Oral health variation mirrored that of many other local populations. Overall prevalence of dental caries in the anterior teeth was 2.2%, and dental caries in the posterior teeth was 15.9%. Antemortem tooth loss was uncommon in the anterior dentition (0.4%) but more common among posterior teeth (4.1%). Calculus was present in 24/31 (or 77.4%) of the people, and its patterning cross-cut all age and sex distinctions. Stable isotope data from the Huaca de los Sacrificios victims reveal their bone carbonate δ 18O values ranged 3.8‰, with a mean of +26.8±1.1‰ (Turner et al. 2013: table 1). Bone δ 13Ccarbonate values ranged 6.1‰, with a mean of –6.7±1.7‰, while bone δ 13Ccollagen values ranged 4.5‰, with a mean of –11.8±1.3‰; bone δ 15Ncollagen values ranged 5.1‰, with a mean of +11.5±1.3‰. Hair δ 13Ckeratin values ranged 5.4‰, with a mean of –12.8±1.6 ‰. Hair δ 15Nkeratin values ranged 5.3‰, with a mean of +10.8±1.3‰. Bone collagen and hair keratin isotope data cannot be compared directly, and differential fractionation must be taken into account (O’Connell et al.

1 9 6   Klaus et al.

2001). In doing so, the estimated δ 13C of dietary sources incorporated into bone collagen for the Huaca de los Sacrificios bone and hair keratin is identical, between –14 and –19‰. Thus, victims at Huaca de los Sacrificios appear to have consumed, on average, a mix of marine and C3 or C4 protein sources both during their final decade and final months of life. The 6‰ range in bone carbonate δ 13C points to moderate variation in their dietary energy sources, with some individuals consuming a mix of C3 and C4 foods and others consuming primarily C4 energy sources. The comparison of collagen to carbonate data suggests some variation in sources of carbohydrate energy compared to dietary protein, but it still falls within the range of values expected in coastal Peruvian ecologies (Turner et al. 2013) and conforms to archaeological expectations (Klaus and Tam 2010; Shimada 1994a). Because there is currently insufficient δ 18Ocarbonate data for northern coastal Peru, a formula-based estimate of δ 18Owater is currently the only feasible method to speak toward residential mobility (Turner et al. 2013:33). δ 18Owater values ranged from –3.1 to –7.1‰, with a mean value of –5.1±1.3‰; (Toyne et al., 2014).

7.12. Composite map of the cut marks documented in all the individuals recovered from Huaca de los Sacrificios. Image by Roy Gutierrez and Haagen Klaus.

Human Sacrifice at Chotuna-Chornancap  197

A

b 7.13. Examples of perimortem sharp force trauma at Huaca de los Sacrificios: (A) cut marks on anterior surface of the C4 and C5 vertebrae, Burial 6; (B) cut marks on the anteriomedial aspect of the left clavicle, Burial 7. Photos by Sam Scholes.

A

b

7.14. Examples of perimortem fractures among the Huaca de los Sacrificios remains: (A) unhealed peri-angular fractures of left ribs 4, 5, and 6, Burial 27; (B) unhealed fracture of the left first rib, Burial 16. Photos by Sam Scholes.

1 9 8   Klaus et al.

Traumatic Injury Evidence of ritual violence at Huaca de los Sacrificios was observed in 29 of the 32 most complete individuals. Unhealed perimortem sharp force trauma was observed on anterior vertebrae for 13/32 (or 40%) of the individuals (figure 7.12). Cut marks clustered around the C1, C2, and C3 regions (figure 7.13a). Wound morphology is again consistent with a metal blade. Throat-slitting and chest-opening co-occurred in 11 victims, and evidence of semidecapitation or decapitation was absent. Evidence of chest mutilation was found in 28/32 (87%) of the skeletons (figure 7.13b, figure 7.14). Variations therein ranged from cuts on the manubrium (5/32, or 16%) to cut marks on the left clavicle (14/32, or 44%), right clavicle (1/31, or 3%), left ribs (13/32, or 41%), and right ribs

7.15. Huaca de los Sacrificios Burial 9, a well-preserved subadult. The mummified contents of the thoracic cavity contain well-preserved lungs and a structure (circled) that appears to be a severed pulmonary trunk. Photos by Sam Scholes.

Human Sacrifice at Chotuna-Chornancap  199

(2/31, or 6%). Additional evidence of chest-opening involved perimortem unhealed rib fractures on the left side of the thoracic cage (12/32, or 38%) and right side of the thoracic cage (6/32, or 19%). Such fractures are consistent with forceful manual opening of the chest cavity. Further, several skeletons at Huaca de los Sacrificios were found with their left ribs outside of normal anatomical position. As decomposition ruptures the costovertebral joint, a rib’s anatomy, gravity, and the downward pressure of the overburden forces ribs to fall inferiorly as the thoracic cage collapses (Duday 2009). In these cases, the right ribs demonstrated textbook collapse, but the left ribs were in vertical or laterally splayed positions. This also commonly co-occurred with perimortem rib fractures and cut marks on the anterior chest wall. In such cases, the left side of the thoracic cage appears swung open and gaping wide at the time of burial. For some time, the motivation behind chest-opening has remained unclear due to a lack of physical evidence—with heart removal, lung removal, collection of more blood, or additional mutilation all possibilities (Klaus et al. 2010). However, Burial 9 (a 9–11-year-old subadult; figure 7.15) was found with the left side of the thoracic cage swung open, and mummified lungs were readily identified. What appears to be the desiccated cross-section of the severed pulmonary trunk (with nothing attached to it) provides physical evidence that chest-opening, at least in this instance, involved heart removal. Mortuary Patterns and Burial Taphonomy The vast majority of victims were wrapped in plain or embroidered cotton shrouds. Body positions in Chambers 1, 2, and 3 (Levels 1 and 2) were quite uniform, with corpses extended on their backs and limbs often flexed. Burial 8 was the only Huaca de Los Sacrificios body placed in a facedown position. The women and children in the north and south compartment were arranged in seated positions with their legs folded underneath them, and some had their arms and hands behind their backs as though bound. All of the Huaca de los Sacrificios victims were positioned to face west. Insect remains (fly puparia, beetle carapaces) were associated with more than 75% of the Huaca de los Sacrificios individuals. The presence of intact and complete fingers and toes indicates interment transpired after at least one blowfly metamorphosis cycle (in this climate, 20 to 30 days), but before the decompositional process began to claim the most

2 00  Klaus et al.

fragile joints (Duday 2009). The opening of Chamber 2 and disarticulation of various individuals followed in situ decomposition. If any connective soft tissue (mummified or otherwise) remained, disarticulation and dragging of semi-articulated elements would have likely occurred. In contrast, seated skeletons in the north and south compartments were quite disarticulated, with mandibles fallen atop their pelvises and thoracic cages collapsed and jumbled. This group appears to have been killed and rapidly sealed into their compartments, where they decomposed in a relatively empty space before sand seeped in. Grave goods were uncommon in the Huaca de Los Sacrificios assemblage. The Chamber 1 subadults were accompanied by seven mummified juvenile camelids, each arranged with their legs flexed and tucked under the belly. These camelids were also likely sacrifices and part of the overall symbolism of Chamber 1, reflective of a “like with like” concept (Gaither et al., chapter 6). One individual (Burial 20) was buried holding a Spondylus shell. A ceramic double-spout vessel in the form of a loche fruit was placed between Burials 9 and 10 and above the head of Burial 13. A pair of undecorated gourds accompanied Burial 21, and part of a drilled shell-bead bracelet had been placed by their right shoulder. Clusters of seeds (preliminarily identified as Gossypium spp., or cotton) were present next to Burials 9 and 20. A maize cob was adjacent to the head of Burial 2. Considering the pristine nature of the Huaca de Los Sacrificios settings, it is very doubtful these materials represent refuse or accidental deposition.

Discussion Diachronic Sequence of Ritual Killings at Chotuna-Chornancap The intra-site sequence, from the Chimú-era killings at Chornancap Norte to the terminal late pre-Hispanic Inka period sacrifices at Huaca de los Sacrificios, highlights ritual violence at Chornancap Norte and Huaca de los Sacrificios at the crossroads of long-standing traditions and the evolution of preexisting themes. Dynamic Rituals Within Conservative Frameworks The deeply patterned elements of ritual killing at Chornancap Norte and Huaca de los Sacrificios link directly to local precedents. It seems that human sacrifice followed the same basic playbook as it had for centuries in the Lambayeque region, with well-defined parallels at Cerro

Human Sacrifice at Chotuna-Chornancap  2 01

Cerrillos and Túcume (Klaus et al. 2010; Toyne 2011a, 2011b). The appropriate range of victims once again involved assortments of children, adolescents, and adults. Victims were afforded relatively proper burials within ceremonial or sacred spaces and interred with few to no grave goods; when present, grave goods included occasional Spondylus shells and camelid offerings, almost never any metal items. Most burials correspond in form to the burial patterns of the long-lived Muchik ethnic group. Yet unlike the standard south-north body orientation, these individuals were positioned east-west to face the sea. Prolonged primary burial and post-depositional alteration of human remains was also an enduring element of funerary rituals in the region (Shimada et al. 2015). Simultaneously, the sacrifices at Chotuna-Chornancap demonstrate change over time. At Chimú-era Chornancap Norte, cut mark distribution demonstrates emphasis on slitting the throat and mutilating the chest. In a few cases, throat-slitting was so extensive or intense that it to semidecapitation. During the Inka occupation, a greater focus on chest mutilation and rib fractures is evident at Huaca de los Sacrificios. Throat-slitting was less common, and semi- or complete decapitation disappeared completely (figure 7.16).

7.16. Comparison of sacrifice-related traumatic injuries at Chornancap Norte (Chimú era) and Huaca de los Sacrificios (Inka era). Image by Haagen Klaus.

2 02   Klaus et al.

Further, at Chornancap Norte, the majority of Chimú era victims with only slit throats were subadults; virtually all the thoracic cut marks were among males, and only one of the women was mutilated. Later, at Huaca de los Sacrificios, chest mutilation spans all age and sex distinctions, and women were extensively mutilated. The horizontal and vertical stratigraphy at Chornancap Norte indicates bodies accumulated in a progressive fashion—that is, an accumulation

7.17a. Age distributions of sacrifice victims at Chornancap Norte and Huaca de los Sacrificios. Image by Haagen Klaus.

7.17b. Sex distributions of sacrifice victims at Chornancap Norte and Huaca de los Sacrificios. Image by Haagen Klaus.

Human Sacrifice at Chotuna-Chornancap  2 03

7.18. Comparison of the crude prevalence of skeletal stress markers, dental caries, and antemortem tooth loss (AMTL) frequencies between Chornancap Norte and Huaca de los Sacrificios. Image by Haagen Klaus.

of individual or multi-individual deaths. At Huaca de los Sacrificios, greater numbers of individuals were evidently killed and buried more or less simultaneously. Here, there was evidence of age- and sex-based spatial segregation. Two males were found together in Chamber 2, and a third male was placed in an adobe brick enclosure in the northeast corner of Chamber 3, Level 2. Virtually all of the other individuals at Huaca de los Sacrificios were placed as clusters of adolescent or adult women with children. Victim Demography, Life Histories, and Identities Both settings included fairly diverse groups of victims, but the distribution of ages follows a skewed pattern (figure 7.17a). Unlike a naturally occurring death assemblage (Coale and Demeny 1966), most

2 04   Klaus et al.

individuals were between 14 and 25 years of age. By the Inka period, the victim average age at death was strongly skewed toward the 5–14 age range, with more than twice the proportion of children. In terms of sex variation, each group contains some combination of adult males and females (figure 7.17b). Here, a key historical detail emerges: during the Chimú era at Chornancap, adult women appear for the first time in a blood sacrifice assemblage in the Lambayeque region. By Inka times at Huaca de los Sacrificios, the sex ratio of sacrifice victims portrays a near-total focus on adolescent and adult female victims. Both groups of victims exhibited similar crude prevalence of biological stress markers (figure 7.18a), and both groups experienced particularly stress-riddled childhoods. This parallels previous findings where poorer developmental health and more cariogenic diets were concentrated among the non-elite, mostly ethnically Muchik social strata in Lambayeque (Klaus et al. 2010; Klaus, in press). Unfortunately, we lack a large or socially representative sample of elite and non-elite Chimú and Inka skeletons for statistical comparison. Still, the most parsimonious and contextual explanation drawn from the current understanding is that the relatively poor childhood and oral health among the victims owes to repetitive selection (a kind of cultural oversampling or targeting) of victims from lower social strata among the region’s population. This conclusion is consistent with a larger body of local data and well-established expectations of health in subaltern social strata in complex societies (Nguyen and Peschard 2003). While detailed studies of ancient DNA and inherited dental traits are still ongoing, there are no tooth morphology variations to suggest the presence of foreign individuals. It is thus notable that the seated/flexed burials at Chornancap Norte have been associated elsewhere with Chimú burial practices (Donnan and Mackey 1978; Klaus and Wester 2005). Perhaps this reflects a local emulation of the Chimú style at Chornancap Norte for reasons that are not yet clear. Further perspective on the identity of the Huaca de los Sacrificios victims is found in δ18Owater data. As discussed elsewhere (Turner et al. 2013), the documented δ18Owater values are fairly diverse, and such values likely reflect consumption of non-identical water sources. Thus, we may infer that victims were likely drawn not from a single, homogenous source but multiple locations or communities. Simultaneously, these values fall directly between δ18Owater values reported to the north and south of the Lambayeque region (Hewitt 2013) and are very similar to isotopic values of local Colonial Muchik peoples from nearby Mórrope. Despite

Human Sacrifice at Chotuna-Chornancap  2 05

the various limitations, it is plausible that victims originated from Lambayeque environs and not distant coastal valleys or the highlands. Belief and Sacrificial Symbolisms The search for meanings in ancient Andean ritual is often challenging, as analogies with ethnohistory or other Andean cultures and times represent a double-edged sword. With such challenges in mind, there are strong associations of these sacrifices with water. Positioning victims to face the ocean, visible from their burial site, probably carried a fairly direct message. Spondylus shells accompanied female victims at both sites. In Late Moche iconography, Cordy-Collins (2001b) and Swenson (2011) consider a conceptual triad linking females, water/Spondylus shells, and sacrificial blood. On the wall below the painted murals at Chornancap, several incised graffiti-like carvings included a Mochicoid Decapitator figure holding a pair of human heads spraying blood (Wester et al. 2010b). At Huaca de Los Sacrificios, three narrow channels were dug out of the architecture, extending downward, literally at the feet of some of the bodies in Chamber 3, Level 1. Even though heavily eroded, they appear analogous to channels at the earlier Huacas Loro, Las Ventanas, and El Corte in the Pomac forest. It is hypothesized that such canals were involved in rituals where liquids were made to flow from the structures, symbolically depicting the huaca as the source of life and water in the world (Klaus and Shimada, chapter 5). Similar canals at Huaca de los Sacrificios in direct association with sacrificed bodies is provocative in terms of potential water symbolisms. Water-linked motifs were also associated with Chimú sacrifices nearby at Túcume, including numerous whole and fragmented Spondylus shells, a carved shell plaque in the shape of a fish, and miniature silver items in the forms of boats, fish, and waves (Toyne 2008). The Rubiños y Andrade (1936 [1782]) manuscript recounts a late pre-Hispanic myth from nearby Mórrope involving the discovery of a life-giving water source in the desert that subsequently received a trio of sacrificed children. Calancha (1972 [1638]) describes Chimú child sacrifice to the moon divinity Si as linked to water, blood, and the ocean (Rowe 1948). Ultimately, these associations share multiple parallels with a variety of Andean settings where the qualities of human blood involved interchangeable metaphors with water as the substance that renewed agricultural and mythic cycles (Arnold and Hastorf 2008; Cieza de Leon 1998 [1553]; Classen 1993; Ramírez 2005).

2 06   Klaus et al.

The killings at Chornancap Norte and Huaca de los Sacrificios may have also embodied fertility symbolisms. A polychrome mural on the courtyard walls that faced the Chornancap Norte victims were decorated with anthropomorphic birds, some of which carried decapitated heads (Donnan 2012; Wester et al. 2010b). Among the birds, at least three frontalfacing humanoid figures are holding plants with leaves, stalks, and roots. Similar plant-like objects emanate from the tops of a few decapitated heads. Heuristically, such motifs are reminiscent of imagery in Paracas embroideries from Peru’s south coast. There, visual metaphors of plant growth were intertwined with wounds, blood, and bodily fluids, such that plants or streamers with bean sprouts emerge from wounds or severed heads that are metaphors for seeds (Frame 2001). Despite the significant gulfs of time, space, and cultures between Paracas society and Chornancap Norte, we wonder if they echoed vaguely comparable concepts. Vegetal metaphors also emerge though the use of grave goods at Huaca de los Sacrificios. Of all the diverse kinds of offerings available, the living chose to place seeds, gourds, and a ceramic bottle in the form of a fruit among the victims. Vegetal metaphors may have extended to the dead bodies. In various Andean traditions, the dead, or even bones themselves, are likened to seeds—the dry, lifeless husks from which new life emerges (Salomon 1995). The burial of dead bodies was probably less akin to “disposal” and more likely represented another offering medium. Practices of infesting the deceased with fly larvae at Huaca de los Sacrificios dates back at least to the Moche (Huchet and Greenberg 2010). Iconographic and ethnohistoric sources suggest this involved an attempt to properly separate body from soul (Bourget 2001b). Perhaps the end result of this process produced desiccated corpses not unlike “life-giving” ancestor mummies, or mallqui, of highland traditions (DiLeonardis and Lau 2004), and sacrificial death was a prerequisite to the regeneration of life. Ritual destruction of life reciprocally enabled creation of life, cosmos, time, and authority, such that these victims were situated at the ultimate nexus of “transforming and becoming” (Swenson 2012:15). It is easy to speculate that focusing on female victims involved beliefs surrounding generative fertility and childbearing. This is probably too simplistic and etic. Still, Swensen and Warner (2012:331) instructively consider the possibility that Late Moche female sacrifices in the Jequetepeque valley literally or symbolically involved procreative and animating

Human Sacrifice at Chotuna-Chornancap  207

powers such as the Andean notion of kamayu, which represented vitalizing and generative deaths sometimes associated with running water. In further assessing female sacrifice, we also need to contemplate other meanings and roles borne by late pre-Hispanic women, as they occupied many powerful and specialized sociopolitical and religious roles on the north coast from Moche times onward (Castillo 2011; Shimada 2009), including an elite woman buried at Chornancap (Wester 2012). Political and Historical Implications Foreign Influences and Conquest Politics Various notions of imperialism and hegemony in the Andes, including that of the Chimú and Inka, were articulated in Kolata’s (2006) models of direct and indirect management of conquered societies. In modes of direct rule, foreign governors of a colonizing state wielded power through coercive force, annexed territory, extracted taxes, resources, and tribute while subordinating local elites and enforcing linguistic and religious change. Over time, subject populations identify, collaborate, and emulate the colonizers as local social identities are replaced (Kolata 2006). Work at Túcume, Pampa de Chaparrí, and Farfán indeed reveal expressions of Chimú and Inka direct rule (Hayashida 2006; Heyerdahl et al. 1995; Mackey 2006, 2010). Still, direct control evidently was not applied to human sacrifice. The basic structural elements of these killings all possess thoroughly local precedents. That is, local people were most likely conducting local rituals and doing so rather autonomously without any obvious direct participation by imperial ruling powers. In a setting where conquerors were concerned with co-opting and controlling local political structures, trade networks, and labor, religion was not a direct focus. We may ponder if seated victim burials at Chornancap Norte may have been less about foreign imposition and more about orthopraxy or strategic mimicry. The same may be said about the presence of the Inka metal figurine at Huaca de los Sacrificios, but it is also equally possible it was a token of Inka participation, sanctification, or benediction. Pre-sacrificial Treatment at Huaca de los Sacrificios A few chroniclers (Acosta 2002; Cobo 1990 [1653]; Molina 1943 [1575]) describe Inka slitting the throats of victims or cutting out their hearts. Reinhard and Ceruti (2010:131–132) make a well-reasoned argument that

2 08   Klaus et al.

such accounts probably conflated Inka practices with those of the Aztecs and that llama heart sacrifices were probably misrepresented as those of human children. Archaeological evidence (Andrushko et al. 2011; Reinhard and Ceruti 2010) demonstrates the pristine, intact nature of Inka victims. Thus, much of what is currently known of sacrifice in Inka state religion contrasts quite strongly with north coast traditions. Yet the possibility of Inka influence at Huaca de los Sacrificios remains. Perhaps Inka administrators in Lambayeque selected who died or guided their preparation for sacrifice within the framework of Inka state religion. For instance, Inka nobility selected sacrifice victims from throughout Tiwantinsuyu on the basis of a set of desirable characteristics. Inka lords used this tactic for selection of yanacona and acllacona retainers to exert power and influence by honoring the members of different client communities (Pease 1982; Rowe 1982), and they also linked sacrifice to political dominance in the same ways. Future supplicants were taken on a pilgrimage from their natal communities to the imperial capital of Cusco, where they were fêted in a sequence of ceremonies and feasts sometimes lasting a year or longer. They were finally sent home or taken to a sacred location, such as a mountaintop, to die. Stable carbon and nitrogen isotopic analyses of hair from mountaintop Inka sacrifice victims bear witness to their preparation and provisioning in a shift from a C3-based diet to one dominated by high-status C4 foods such as maize and terrestrial meat during the months leading up to their deaths (Wilson et al. 2007). Isotopically, there is little to no variation in dietary composition between bone and hair tissues at Huaca de los Sacrificios. They do not appear provisioned with high-status or uncommon foods for any prolonged period before death. Had they been taken from Lambayeque to Cusco and fêted, dietary change should have registered in their tissue chemistry. They do not appear to have been prepared in accordance with known Inka rituals (Turner et al. 2013), further suggesting the Inka were not direct participants in ritual violence at Chotuna. Still, one element that could reflect concrete Inka ideas involved hair-offering. Most of the adolescent and adult women at Huaca de los Sacrificios had preserved hair, but their hair was markedly short, and in a few cases, hair was so well preserved that it could be determined that it was clearly shorn in the perimortem period (figure 7.19). In its autonomous ability to grow, hair may have shared bridging metaphors with sacrificial blood, representing “the condensed energy of the head, flowing

Human Sacrifice at Chotuna-Chornancap  2 09

7.19. Huaca de los Sacrificios Burial 24. This woman’s hair appears to have been cut during the perimortem interval. We suspect hair may have been another sacrificial medium. Photo by Sam Scholes.

out from it in a voluptuous expression of vitality, and perhaps, fertility” (Weismantel 2015:89). For the Inka, hair was a sanctified offering (Reinhard and Ceruti 2010:158–159) involving multifaceted meanings of life force. Inka capacocha sacrifices, and the mountaintop platforms they were buried in, both received human hair offerings. In three cases atop Llullaillaco, victims received offerings of their own hair, which had been cut about six months before death (Wilson et al. 2007). The ultimate fate of the cut hair at Huaca de los Sacrificios is unknown. Connections with Inka ideas remain an intriguing possibility.

Conclusion Much of the Chotuna-Chornancap complex remains unexplored under sand dunes. Future excavations atop Huaca Chotuna, at the adjacent Huaca Susy, and in various plazas and platforms will certainly expand current understandings. Pedestrian survey alone has identified disturbed skeletal remains of multiple subadults atop the heavily looted summit of Huaca Chotuna, and what is described here probably reflects only part of a larger ritual program at this site. Integrated studies of ancient DNA, dental phenotype, and proteomic variation are currently under way to

2 1 0  Klaus et al.

assess genetic and social relationships among the victims. Contextual studies of sacrifice victims provide perspectives for reconstructing the wider society at the end of the pre-Hispanic era. The emerging picture is complex, with evolving practices and permutations of well-established local rituals involving symbolisms of sacrifice, water, and fertility weathering two waves of foreign rule. Local people seemingly carried out their most sacrosanct religious acts with autonomy, seeking to maintain the delicate balances between the forces of life and death until European contact brought the pre-Hispanic epoch to its end. Acknowledgments The generous support of the Ministerio de Culture’s Unidad Ejecutora 005 Naymlap-Lambayeque has funded the excavations at the Chotuna-Chornancap Archaeological Complex since 2006. Utah Valley University’s Office of Engaged Learning and College of Humanities and Social Sciences supported excavation and analysis of the human remains by Haagen Klaus and students between 2008–2011. We are especially indebted to Scott Applegate, Jerel Bartholomew, Sylvia Bentley, Elizabeth Byrnes, Joseph Luce, JoEllen Perez, Kat Phillpotts, Sam Scholes, Becky Ann Talpas, and Justin Hadley-Nelson who ably assisted in the field and lab work collection from 2009–2011. We thank Clark Larsen, Go Matsumoto, J. Marla Toyne, and four anonymous reviewers for critical comments on earlier versions of this chapter.

Life Before Death  2 1 5

chapter eight

Life Before Death A Paleopathological Examination of Human Sacrifice at the Templo de la Piedra Sagrada, Túcume, Peru J. Marla Toyne Department of Anthropology, University of Central Florida, Orlando

Who were the victims of the Temple Sacrifice at Túcume? This straightforward question implicates one of the most important and complex lines of inquiry for the global archaeology of ancient ritual killing. Understanding who was being selected for sacrifice deeply informs the reconstruction of the historical and social significance underlying acts of ritual violence. Frequently, the culturally diagnostic mortuary features so often relied upon to infer social identity in Andean archaeology—such as grave goods, adornments, tomb construction, and group-specific cemetery settings—are absent. Accordingly, this approach to examining ancient identity challenges bioarchaeological methods, forcing the development of alternative and often highly nuanced means of assessing the identity of sacrifice victims. As human skeletal remains can provide a distinct record of an individual’s lived experience prior to death, the analysis of skeletal pathological conditions may aid in the reconstruction of aspects of social identity, particularly in complex societies where social organization directly shapes elements of human biological realities. In this chapter, I focus less on patterns of skeletal trauma and emphasize exploring victim identity among the sacrifices from the Templo de la Piedra Sagrada (Temple of the Sacred Stone, or TPS) during the Chimú (AD 1350–1470) and Inka periods (AD 1470–1532) at Túcume.

2 1 6   Toyne

Demographic, paleopathological, and body modification data are used to reconstruct the social identity of these victims through the creation of a biological health profile, or the osteobiography of individuals, who are then assessed together as a group. Comparisons are made with other, non-sacrificial burials from Túcume and other samples from farther afield to elucidate who the victims were, where they may have come from, and why they were selected to die in ritual. Comparisons with regional archaeological contexts and ethnohistoric accounts of Inka human sacrifice and perimortem treatment further inform paleopathological approaches toward victim identity at Túcume.

The Templo de la Piedra Sagrada: Site and Contextual Description Túcume is located in the La Leche river valley, part of the greater Lambayeque River Valley Complex on the northern coast of Peru. Encompassing more than 200 hectares, Túcume’s monumental core is centralized around the prominent mountainous rock outcropping of Cerro La Raya. The principal structures included 26 massive mud-brick platform mounds serving administrative, economic, and ritual functions. Initially settled around AD 1000 by the Lambayeque (or Middle Sicán) peoples, the site was later conquered by the Chimú Empire (ca. AD 1350) as it expanded northward along the coast. Later, Túcume was incorporated by the Inka (ca. AD 1470) (Narváez 1995b). During these eras of foreign rule, Túcume was the largest paramount regional administrative center for the Chimú and Inka (Heyerdahl et al. 1995; Shimada 2000). Site chronology was established through long-term excavations in both ceremonial and urban sectors, where the use of radiocarbon dates and ceramic seriation document these successive cultural occupations. Each occupation is clearly identified through defined shifts in material culture and substantial modifications to existing architecture reflecting the transition from the preexisting sociopolitical order. Yet continued use of local symbols and sacred spaces suggests an important level of continuity. The major cemetery areas at Túcume were thoroughly destroyed over centuries of looting; however, the excavated burials from the Southern Cemetery date from the earliest occupation of the site to the Early Colonial period (Narváez 1995a). These mortuary contexts demonstrate a fairly consistent funerary program involving extended body positioning (oriented head to the north and feet to the south) or seated positions with the legs flexed, especially during Chimú and Inka times. Individuals

Life Before Death  2 17

were accompanied by a variety of material offerings, including ceramics, weaving implements, and personal adornments (Narváez 1995a). The range of elaboration within these burials is consistent with middle-level social status based on access to specialty trade items, such as Spondylus and small objects of bronze and copper. During both Chimú and Inka rule, many of the earlier Lambayeque iconographic and stylistic traditions continued to be expressed in grave associations and rather unambiguously suggest continuity of local traditions and peoples (Narváez 1995a; 2005). At other sectors at Túcume, including atop Huaca Larga, Huaca I, and Huaca Las Balsas (Narváez 1995a; Narváez and Delgado 2011; Narváez and Delgado, in press), recently excavated burials provide excellent evidence of other mid-tier social status burials. None of these burials can be characterized as social elites, especially when compared to tombs such as those at Sipán or Sicán. Indeed, most burials at Tú-cume involve simple pits and lack ostentatious quantities or accompanying retainer burials. A small temple designated as the Templo de la Piedra Sagrada is located at the east base of Huaca Larga, which was the largest monumental construction at Túcume at over 700 meters long. The temple was constructed around a vertically implanted stone monolith during the early Late Intermediate Period (ca. AD 1100) and appears to have been a small but sacred center at Túcume. At least five construction phases were identified involving the addition of walls, benches, and a doorway to the structure (Narváez 1995b). Within the small enclosed patio to the north of the temple entrance, a large number of human and camelid remains were found within individual burial pits (Narváez 1995b; Toyne 2008) (figure 8.1). Almost all of these individuals (98%) were oriented with their head to the east and feet to the west. Although the bodies were supine, the positioning of the limbs was variable, either extended or flexed and crossed. Such positioning is consistent with mortuary contexts at Túcume and might not be considered distinctive (Toyne 2008) (figure 8.2). However, grave goods were completely absent other than remains of poorly preserved simple cotton shrouds that wrapped the bodies. At least 83 juvenile camelid (llama) remains accompanied these burials and had evidence of throat-slitting but had not been dismembered or defleshed for food. Each of these was either buried in a separate pit or associated with another camelid or human body. Most of the human skeletal remains (94%) demonstrated evidence of perimortem sharp force trauma to the upper thorax and neck consistent

2 1 8   Toyne

8.1. Photo of Túcume archaeological structures and Templo de la Piedra Sagrada (center). Photo by J. M. Toyne.

8.2. Sacrifice victim Burial 85, demonstrating the degree of completeness and excellent preservation at the Templo de la Piedra Sagrada. Photo by J. M. Toyne.

Life Before Death  2 1 9

with having their throats slit, heads removed, chest cavities cut open, or, more frequently, a combination of these activities (figure 8.3; Toyne 2011b). Clearly, these individuals were subjected to a complex sacrifice ceremony involving the collection of blood, removal of the heart, and decapitation. These rituals were performed on a regular and possibly frequent basis, perhaps timed to specific calendrical events (Toyne 2008, 2015a). Discerning an exact chronology of the Temple burials is complicated by complex superposition of individuals, numerous remodeled floors, reuse of pits, and floor erosion. Still, the consistent pattern in the remains reflects a sequential accumulation of bodies spanning the

A

b

8.3. Sharp force trauma on the skeletal remains of Burial 5 including cuts on (A) anterior c2 vertebrae and (B) a bisected medial clavicle. Photo by J. M. Toyne.

2 2 0  Toyne

Late Intermediate Period through the Late Horizon (AD 1300–1532). Even though major changes in political and religious administration of the site occurred atop the adjacent Huaca Larga, this small temple continued to be used for the burial of ritually killed people and camelids. Human offerings buried in the Temple patio continued unchanged at the site during the period of Inka control, demonstrating a strong continuity of local traditions and ritual beliefs mirroring the nearby ChotunaChornancap archaeological complex (Klaus et al., chapter 7). The sacrifices at TPS could be consistent with two distinct models of sacrifice victim identity. The identical pattern in perimortem violence and burial treatment suggests each person experienced a similar ritual program, perhaps linked to some shared aspects of social identity and ritual practice. Alternatively, the systematic treatment may suggest a common ritual template regardless of victim identity and that these individuals may have diverse social origins relatively unrelated to the sacrificial program. To test these two possibilities, multiple, independent, and complementary skeletal indicators of lived experiences and health can be utilized to reconstruct aspects of antemortem individual lives. This research characterizes victim demography, nonspecific indicators of childhood stress (cribra orbitalia, porotic hyperostosis, and linear enamel defects), adult stature, patterns of antemortem trauma, and cranial modification recorded in individual skeletons to create osteobiographic profiles of those interred at the Templo de la Piedra Sagrada. Specifically, I hypothesize that the victims will demonstrate withingroup homogeneity across skeletal health indicators, consistent with shared group experiences and a single origin. When compared to a non-sacrificial cemetery from Túcume, I argue that sacrifice victims will differ from the inferred local population. Additionally, as may be the case with the earlier Moche (e.g., Sutter and Verano 2007; Toyne et al. 2014), Túcume victims may have come from distinctly different or distant locales, lived different kinds of lives, and may have represented different social standings than the local people interred at Túcume.

Skeletal Indicators of Physiological Stress and Lifestyle An embodied perspective on past population biology proposes that physiological indicators of stress and lifestyle, including daily activities and health experiences, are variably expressed in human skeletal remains and reflect lived social experiences (Robb et al. 2001; Sofaer 2006). The

Life Before Death  2 2 1

osteobiographic approach holds that bones can reveal elements of individual life histories, making it possible to reconstruct various kinds of life experiences (Saul and Saul 1989). For example, an ever-growing body of work in health theory, bioarchaeology, and epidemiology shows that social status within complex societies shapes health and other aspects of biology (among many others, e.g., see Danforth 1999; Klaus 2014c; Larsen 2015; Nguyen and Peschard 2003). Additionally, individuals sharing similar roles in society tend to have access to similar resources, quantities of socioeconomic capital, and cultural identities that more likely to create shared patterns of biological health status. Thus, combined osteobiographies create a collective profile, although there may be individual idiosyncrasies or variations in frailty in these outcomes. Bioarchaeologists recognize that interpretations of skeletal phenomenon are not straightforward. For instance, not all biological stressors affect the skeleton, but those that do are often consistent with prolonged or chronic health conditions deeply embedded within lived experience. Issues surrounding the “osteological paradox,” especially in terms of sample variability in “frailty” and “hidden heterogeneity” (Wood et al. 1992; Wright and Yoder 2003), are also relevant. Yet variation in social status is known to directly shape health over the life course involving overarching socioeconomic structures that influence access to nutritional resources, medical treatment, stress levels, and workload. Thus, these phenomena relate to the frequency of events that leave skeletal indicators (Goodman and Martin 2002; Klaus 2012; Robb et al. 2001). Poorer health and greater stress appear to be most common among individuals of lower social status (Larsen 2015; Nguyen and Peschard 2003; Salpolsky 1994). Specific skeletal indicators (see below) can help us understand some of the complex manners in which social status is reflected in human biology (Bush 1991; Goodman and Martin 2002). Therefore, the combination of multiple lines of contextual archaeological data and skeletal pathological conditions can elucidate a general understanding of individuals’ social position in complex societies. Paleopathological Data and Patterns of Health in Ancient Andean Societies Research around the world demonstrates that paleopathological data play an important role in understanding social differences between individuals found in traditional and nontraditional cemetery collections, including defining possible sacrifices (Buzon and Judd 2008; Coughlan

2 2 2   Toyne

and Holst 2000; Tiesler and Cucina 2007). Previous studies of Andean skeletal samples from chronologically and geographically distinct archaeological sites have used skeletal pathology to characterize what life was like for populations at specific sites or regions (Blom et al. 2005; Klaus et al. 2009; Klaus and Tam 2009b; Toyne 2006; Tung 2012; Tung and Castillo 2005; Verano 1986a, 1992, 1997c). Some researchers have focused on specific aspects of ancient population health, including dietary variation (Benfer 1990; Kellner and Schoeninger 2008; Lambert et al. 2012). Others have explored relationships between social differentiation and relative health status and morbidity using multiple skeletal features in various regional cultures (Andrushko et al. 2006; Farnum 2002; Klaus et al., in press; Murphy 2004; Pechenkina and Delgado 2006; Yoshida 2004). A coherent picture of the prevalence of morbidity markers is still emerging in Andean bioarchaeology, and more research is required from different regions and time periods. Still, the skeletal health outcomes of elite individuals on the north coast, while complex, tend to demonstrate lower frequencies of many pathological conditions (Klaus et al., in press; Shimada et al. 2004; Verano 1997b). Artificial cranial deformation is a characteristic that has been used successfully in the southern region of the Andes to differentiate social origins of individuals (Blom 2005; Lozada and Buikstra 2005). However, a comprehensive or chronologically well defined study of Peru’s northern coast has yet to emerge, prohibiting similar evaluations of the potential relationship between cultural modification of the skull and social identity. Work by Klaus and colleagues (2010) at Cerro Cerrillos in the southern Lambayeque region identified variation in skeletal health indicators among sacrificed individuals, suggesting the victims may have originated from the lower social strata of the local population. Further work has considered broader regional and diachronic studies from a spectrum of mortuary contexts. These findings suggest rather rigid, socially structured variability in mortuary treatment and that lower status correlated with higher levels of biological stress, especially in the late pre-Hispanic and Colonial periods (Klaus and Tam 2009a, 2009b; Klaus et al., in press). Klaus and colleagues (chapter 7) also observe that high levels of biological stress characterize sacrifice victims at the nearby ChotunaChornancap archaeological complex. In examples outside the Andes, Tiesler and Cucina (2007) explored identity and health among the ancient Maya of Mesoamerica. They

Life Before Death  2 23

identified greater skeletal stress among individuals in non-funerary contexts, including sacrifices, and indicate these individuals might have come from lower social status groups. Buzon and Judd (2008) compared apparently sacrificed juvenile and adult individuals and a contemporary cemetery group expecting different health profiles, but they found a lack of observed differences. Such results might be obscured by a number of factors, including archaeological misinterpretation of sacrifice and the lack of children for comparison, as many of the indicators observed in adults reflect survived childhood experiences.

Materials and Methods A total of 132 individuals were documented in excavations at the Templo de la Piedra Sagrada. Most were well-preserved and complete skeletons, although some burial pits contained incomplete skeletons or isolated crania (n = 29; 24.8%). The long history of use and reuse of the ritual space has spanned approximately AD 1300 to AD 1532 (Toyne 2008). Non-sacrificial burials excavated from a variety of funerary contexts at Túcume represented a slightly broader time frame (AD 1100–1532) and included 137 individuals. A few of these burials are interpreted as possible sacrificial offerings or secondary dedicatory remains (n = 19), but these were all females interred with similar quantities and qualities of finely crafted weaving implements, suggesting a consistent mid-level social standing (Toyne 2002). None of these individuals demonstrated evidence of violent death, and for this study they were considered as non-sacrificial contexts. These two groups at Túcume are the basis for evaluating relative health between these different sacrificial and nonsacrificial contexts. Several methods were used to document skeletal health, beginning with observations during excavation and later during comprehensive laboratory study. Age and sex were estimated using standardized methods (Buikstra and Ubelaker 1994). Age of subadults was estimated based on dental eruption, long bone epiphyseal union, and long bone length (Moorees et al. 1961; Schaefer and Black 2007; Ubelaker 1989). Adult age was estimated from the degeneration of pubic symphysis and auricular surface morphology. In the absence of os coxae, age was approximated using a combination of dental attrition, articular degeneration, and degree of cranial suture closure (Lovejoy et al. 1985). All pathological conditions were recorded via detailed descriptions of the location,

2 2 4   Toyne

extent and degree of affliction, and evidence of healing. While crude prevalence of presence/absence data of stress markers has been critiqued as potentially misrepresenting patterns of stress (Waldron 1994), similar sample sizes and structures here do allow for direct comparisons between sacrificial and non-sacrificial groups and with other published datasets. Statistical comparisons were made using Fisher’s Exact test for smaller sample sizes, and significance was assessed at p = < 0.05. Cribra Orbitalia and Porotic Hyperostosis In childhood, various kinds of chronic anemia produce lesions involving increased thickness and coalescing porosity of the outer table of the cranium, as erythropoiesis increases the volume of interdiploic marrow (Stuart-Macadam 1992; Ortner 2003; Walker et al. 2009). Cribra orbitalia occurs on the roofs of the orbits, while porotic hyperostosis typically affects the posterior cranial vault. Traditionally, both lesions were believed to relate to iron-deficiency anemia and maize-dependent diets (Stuart-Macadam 1992). More recently, others have argued that skeletal anemia is more likely due to variable nutritional insufficiencies (Walker et al. 2009), parasite load (Tung 2005), infectious disease burdens (Blom et al. 2005), and settlement patterns (Klaus and Tam 2009b). While a specific etiology cannot be determined, such lesions indicate chronic systemic stress affecting hematopoietic tissue during childhood. Enamel Defects During growth and development, accretional deposition of tooth enamel can be disrupted in response to various types of physiological stress (Goodman and Armelagos 1985; Goodman and Rose 1991). Hypoplastic enamel defects (linear and pits) are most commonly observed on the labial surfaces of anterior teeth and represent a critical bioarchaeological window on childhood metabolic stress reflecting the slowdown and then the re-initiation of normal growth after a stress episode. Adult Stature Estimation As growth is a highly sensitive measure of health and socioeconomic status (Tanner and Eveleth 1990), achieved adult stature also reflects aspects of childhood health and nutrition. Measurements of femoral

Life Before Death  2 2 5

and tibial long bone lengths to estimate adult stature (del Angel and Cisneros 2004; Genovés 1967) can be compared among samples along the north coast with the expectation that significantly reduced stature reflects some degree of insufficiency or stress that inhibited full growth potential (Schumacher 1982; Steckel 1995). Antemortem Traumatic Injuries Antemortem trauma (fractures, dislocations, or ossification of ligaments due to muscular injuries) can demonstrate aspects of lifeways and physical activities (Larsen 2015; Lovell 1997). Anthropologists broadly divide traumatic injuries into two categories: (1) violent origin (interpersonal trauma) and (2) nonviolent origin (accidental, occupational, pathologyrelated, or environmental) (Judd 2002). In some cases, however, differentiating these categories is challenging. A high frequency of cranial and facial trauma (blows to the face and head by an attacker), direct force ulnar fractures (defensive wounds), and the presence of multiple injuries in the same individual are likely associated with combat situations or violent interpersonal interactions (Brickley and Smith 2006; Judd 2004; Walker 2001). Nonviolence-related injuries are assumed to reflect occupational or environmental hazards and are distributed more randomly throughout the skeleton. Cranial Modification Cranial modification in ancient Peru involved the artificial flattening or elongating of the vault through the application of specialized devices (Blom 2005; Imbelloni 1933). Unfortunately, social or temporal practices of cranial modification have not been well defined on the north coast, and currently little published comparative data are available to suggest a correlation with elevated status (but see Klaus [2008a]; Shimada et al. [2004]; Verano [1997a, 1997b, 1997c]). Verano (1997c) suggested that child-care practices involving cradle-boarding were inadvertently responsible for the subtle flattening observed among many pre-Hispanic samples, but there are some examples of extreme forms of intentional modification associated with individuals of inferred higher social status (Nelson et al. 2000), including at Túcume (Toyne 2002) and Sicán (Farnum 2002). In this work, cranial modification was observed as present or absent, by the type of flattening (frontal, occipital,

2 2 6   Toyne

or annular), and by the degree of modification (minor, moderate, or pronounced deformation).

Results Demography The analysis of the sacrifice sample resulted in a minimum number of individuals (MNI) of 116.1 There were 69 adults and 47 juveniles (< 18 years) in the sample with no significant difference in adult:juvenile distribution (χ2 = 2.602, df = 1, p = < 0.10). Average age at death ranged from approximately 4 through 45 years of age, with the largest age group (43%) being young adults (19–34 years) (table 8.1). All adult remains were male, and many late adolescent individuals demonstrated morphological characteristics consistent with males. One female was buried to the side of the temple entrance, perhaps part of a temple dedication or foundation sacrifice, as her remains did not demonstrate cut marks. The skeleton of another middle-aged female was piled as a secondary offering on top of a young adult male who was also secondarily bundled in Table 8.1. Demographic Distribution of the Sacrificial Sample from Templo de la Piedra Sagrada and the Túcume Cemetery Age Category

TPS Sacrifices Male Indeterminate Female

Túcume Cemeteries Male Indeterminate Female

0–11 months



0 – –

1–2 years



0 – – 13 –

2–5 years



1 – –

8 –

6–10 years



22 – –

5 –

11–14 years

0

19 0 0

3

2

15–19 years

9 0

0

0

2

2

20–34 years

59 0

1

11

0

53

35–49 years

18 0

1

7

1

10

0 0

0

1

0

6

50+ years

4 –

Subadult undetermined –

0 – –

2 –

Adult undetermined 0

2 0 3

1

Total

86 44

2 22

3

37 76

Life Before Death  2 27

A

b

8.4. Comparative mortality profiles between sacrifice victims at Templo de la Piedra Sagrada and the burials in the Túcume cemetery: (A) comparative sex distribution and (B) comparative age estimate distribution. Image by J. M. Toyne.

elaborate polychrome textiles. These demographic patterns are wholly different from a normal attritional cemetery, where skeletons representing infants through older adults and both sexes are expected (Bishop and Knüsel 2005; Margerison and Knüsel 2002). The sacrifice sample age distribution is also distinct (figure 8.4a, figure 8.4b) when compared to the cemetery sample at Túcume. In both groups, young adults make up the largest portion of individuals. While

2 2A8   Toyne

b

8.5. Photo of cribra orbitalia and porotic hyperostosis: (A) Individual E6 with active minor porosity of superior orbit; and (B) E21 with active minor porosity of posterior vault. Photos by J. M. Toyne.

there is a similar distribution of juveniles to adults in both groups with a ratio of approximately 1:2, there is a significant difference in age distribution (χ2 = 50.63, df = 1, p = 0.001). Most noticeably in the younger age categories, the cemetery sample includes newborns and infants, and the sacrifice sample includes a higher concentration of adolescents. The sacrifice sample also lacks older individuals (greater than 45–50 years). Sex distribution demonstrates another significant difference; while the majority of the sacrifice samples were males, the cemetery sample at Túcume includes a larger proportion of adult females (79% of the adults). This discrepancy represents a potential shortcoming in comparing the pathological profiles of these groups, since sex-based physiological responses to stress are well documented along with gender-based behavioral differences (Cohen and Bennett 1993; Robb 1997). Similar issues could be raised with the inability to estimate non-adult sex, as the same factors could impact the relative health status of juveniles. Regardless, it remains possible to compare specific differences between juvenile and adult groups. Childhood Anemia Cribra orbitalia was present in 10 juveniles and 4 adults (or 14.4%) with observable orbits, and 28% of these cases demonstrated healing. Porotic

Life Before Death  2 2 9

hyperostosis was observed in 11 juveniles and 24 adults (or 36.1%) with observable vaults, and nearly 60% of these cases were inactive and healed cases. Within the sacrifice sample, juveniles were more likely to have cribra orbitalia (Fisher’s Exact, p = 0.018) (figure 8.5a, figure 8.5b, and figure 8.6). Of the affected crania, lesions were qualitatively minor. A comparative view of these pathological conditions between the sacrifice sample and the cemetery sample (table 8.2) shows that while A

b

8.6. Comparisons of cribra orbitalia (A) and porotic hyperostosis (B) between sacrifice victims at the Templo de la Piedra Sagrada and the burials in the Túcume cemetery.

23 0  Toyne Table 8.2. Summary of Biological Stress Indicators of TPS and Comparison to Other Regional Samples  

 

 

 

Sacrifice

Non-Sacrifice

Subadult

10/40 (25.0%)

2/35 (5.7%)

6/17 (35.3%) 14/29 (48.3%) 3/24 (12.5%)

(presence/absence)

Adult (male)

4/57 (7.0%)

1/18 (5.6%)

3/8 (37.5%)

Porotic Hyperostosis (presence/absence)

Subadult 11/40 (27.5%) 9/35 (25.7%) 12/17 (70.6%) 11/29 (36.7%) 3/24 (12.5%) Adult (male) 24/57 (42.1%) 3/18 (16.7%) 3/8 (37.5%) 12/84 (14.3%) 44/78 (56.4%)

Enamel Defects (presence/absence)

Subadult 33/41 (80.5%) 4/28 (14.3%) 2/16 (12.5%) Adult (male) 44/57 (77.2%) 8/13 (61.5%) 4/10 (40.0%)

Cribra Orbitalia

Túcume

Traumatic Injuries Subadult 4/47 (8.5%) 0% (number of individuals) Adult (male) 27/69 (39.1%) 5/21 (23.8%)

 

 

Farfan

a

1/17 (5.9%) 7/8 (87.5%)

 

Pacatnamú

Punta Lobosd

b, c

17/84 (20.2%) 11/78 (14.4%)

no data no data

no data no data

no data no data

1/47 (2.1%) 46/133 (34.6%)

Adult Stature Estimate Adult (male) 159.14 (n = 47) 159.94 (n = 9) 161.33 (n = 8) 157.8 (n = 17) 160.97 (n = 82)   Cranial Modification (presence/absence) a

(152.19–165.19) (155.1–164.24) (158.14–165.65) (148.2–164.6) (151.80–168.75)

(range)

Toyne (2006)

b

Subadults 9/36 (25.0%) 16/35 (45.7%) 12/13 (92.3%) 0% 14/25 (56.0%) Adult (male) 19/57 (33.3%) 32/73 (43.8%) 8/8 (100%) 38/96 (40.0%) 47/75 (63.5%)

Verano (1997b)

c

Includes both adult males and females

d

Verano and Toyne (2011)

neither group differed in the overall rate of orbital lesions, adult male sacrifice victims were more frequently affected by porotic hyperostosis. Contrastingly, sacrificed juveniles demonstrated a higher rate of cribra orbitalia (half active and half healed cases), but no difference in porotic hyperostosis existed among juveniles outside the sacrificial setting at Túcume. Both indicators suggest that, among the sacrifices, the children experienced higher rates of anemic stress leading to orbital lesions, and the adults had survived more of the earlier life-stress episodes that left vault lesions. The sacrifice victims demonstrated higher levels of anemiarelated physiological stress when compared to the cemetery burials. Childhood Metabolic Disruption The excellent preservation of dental tissues allowed for observation of anterior tooth crowns of 57 adults and 41 subadults from the TPS. Defects were observed in 77 individuals (or 78.5%). Adults demonstrated

Life Before Death  231

a shared pattern of hypoplastic defects with juveniles (Fisher’s Exact, p = 0.6150). Comparatively, those in Túcume’s cemetery demonstrate a significant difference between juvenile and adult groups, where the local children have fewer enamel defects. Growth and Adult Stature A total of 47 adults from the TPS and 9 adult males from the rest of the Túcume sample had complete femora or tibia for measurements, resulting in an average of 159.1 cm ± 3.13 and 159.9 cm ± 3.0 respectively. These data demonstrate no significant difference. Antemortem Traumatic Injuries The TPS injury profile demonstrates both interpersonal and accidental traumatic injuries. Of those sacrificed at the TPS, 31 possessed antemortem traumatic injuries totaling 47 separate affected elements, including four dislocation injuries (table 8.3). Adults (n = 27) and juveniles (n = 4) had broken bones, but juveniles significantly less so (Fisher’s Exact, p = 0.005). Almost all injuries demonstrated evidence of advanced, longterm healing. The one exception was Sacrifice Entierro 29 (30–34 years old), who suffered a perimortem impact to the frontal bone superior to the left orbit. This isolated case was the only example of a perimortem insult, perhaps related to sacrifice. Blunt force cranial trauma was a ritual technique the earlier Moche used to dispatch victims (Verano 2001b). Traumatic injuries possibly associated with interpersonal violence included depressed cranial fractures and nasal bone fractures. Only five individuals had multiple injuries—two adults with nasal and rib fractures (possible interpersonal origin), a young child (with cranial and humeral fractures), and an adult (with cranial and toe fractures). The other individual with multiple injuries was a young adult male (Sacrifice Entierro 5) who recovered from a complex “pilon” fracture of both the right distal tibia and fibula. This individual also possessed a healed Colle’s fracture of the right distal radius that, based on the degree of healing and sidedness, could have occurred at the same time as the fractures in the lower limb. These injuries were classified as accidental, as they are most consistent with a major fall. Other accidental healed injuries included a shoulder injury where the humeral head was crushed at the anatomical neck (figure 8.7), as well as

23 2   Toyne Table 8.3. Distribution of Antemortem Skeletal Fractures by Skeletal Element

Adults

Subadults

Adults & Subadults Combined

Skeletal Element

N

N

N

N

%

Cranial vault

5

2

7

97

7.2%

Nasal region

8

0

8

97

8.2%

Total Cranio-Facial Trauma

13

2

15

97

15.5%

 

N observable skeletal elements

Percent of Cranio-Facial Trauma

 

34.9%

Scapula

1

0

1

182

0.5%

Humerus

1

1

2

187

1.1%

Radius

2

0

2

179

1.1%

Ulna

0

0

0

183

0.0%

Metacarpals

0

0

0

181

0.0%

Femur

0

1

1

192

0.5%

Tibia

1

0

1

178

0.6%

Fibula

1

0

1

183

0.5%

Metatarsals

1

0

1

177

0.6%

Ribs

13

0

13

2161

0.6%

Vertebral Fractures

4

0

4

2208

0.2%

Sacrum

1

0

1

100

1.0%

1

0

1

88

1.1%

26

2

28

6017

0.5%

Sternum Total Post Cranial Trauma  

 

Total Trauma

39

4

Percent of Postcranial Trauma

Joint Injuries

2

2

43

6114

65.1% 0.7%

other long bone fractures, hip and elbow dislocations, and vertebral fractures. Sacrifice Entierro 64 (18–22 years) experienced a hip dislocation of the left leg of either traumatic or developmental origin, resulting in pronounced joint degeneration and deformation of the acetabulum. Evidence of atrophy of the left femur (reduction in diameter and robusticity) and hypertrophy of the right humerus suggest this individual probably ambulated using a compensatory aid, such as a crutch or walking stick. Ribs were the most frequently fractured skeletal element, followed by nasal bones and cranial fractures. Craniofacial fracture rates appear high (22.8% of adult crania). However, these injuries were neither extensive

Life Before Death  23 3

nor associated with other healed fractures. Five of the seven depressed vault fractures were small, button-sized depressions (figure 8.8). A chi-square analysis demonstrates a statistically significant difference between cranial and postcranial injuries, where more injuries were present in postcranial elements (χ2 = 4.454, df = 1, p = 0.02). Most of the postcranial injuries appear related to physically active lifestyles, with the majority found in adolescents and younger adults (between 20 and 25 years). The older adult individuals demonstrated two Colle’s fractures of the wrist and four lumbar vertebrae affected by spondylolysis, most possibly related to falls or other strenuous activities (Merbs 2002). One child (Sacrifice Entierro 12, between 9–11 years old) suffered a compound fracture of the left femoral diaphysis, which was in the process of healing at the time of death (figure 8.9). This injury reduced the femoral length and involved a moderate degree of displacement (twisting), which would have resulted in a significant change in gait, most likely producing a noticeable limp. Compared to the Túcume cemetery sample, the individuals at the TPS demonstrated a much higher frequency of traumatic injuries, suggesting a more physically demanding lifestyle on average. In comparing

8.7. Burial E28 displayed a well-healed crush fracture of the right humeral head compared to the unaffected left side. Photo by J. M. Toyne.

23 4   Toyne

8.8. Cranial trauma (TPS Burial E79, subadult, with small depressed healed fracture of central frontal bone). Photo by J. M. Toyne.

8.9. Fracture of the femur (TPS Burial E12, subadult, with healing complex fracture of midshaft of right femur with significant misalignment). Photo by J. M. Toyne.

only adult males, trauma rates for sacrificed individuals (27/69; 39.1%) are nearly double that of the adult males in the cemetery sample (5/21; 23.8%) with the exception of craniofacial trauma, which demonstrates an almost identical frequency in both groups. Many limb injuries in the sacrificial group are likely related to accidental or occupational trauma and would have been physically noticeable in the living individual, perhaps impeding mobility or economic productivity.

Life Before Death  23 5

Cranial Modification Only 30.1% of sacrificed individuals demonstrate cranial modification, indicating it was not a universal cultural practice (table 8.2 and table 8.4). At Túcume, cranial modification was variable, from minor deformation to pronounced flattening of the frontal bone or the occipital bone (generally near lambda) creating a tabular erect style (Imbelloni 1933; Lozada and Buikstra 2002). In pronounced cases, the bilobate lateral bulging of the parietal bosses would have likely been noticeable in life (figure 8.10a, figure 8.10c; compare with unmodified crania in figure 8.10b, figure 8.10d). However, many cases of minor and asymmetrical flattening of the posterior aspect of the cranium may have been unintentional and are consistent with cradleboard usage. Comparisons at Túcume demonstrate no significant difference in cranial alteration between sacrifices and cemetery burials, regardless of age, modification type (fronto-occipital or just occipital flattening of the vault), or degree of vault modification (minor, moderate, or pronounced flattening). These data suggest a similar range of variation in the cultural practice of cranial modification, regardless of whether it was an intentional act or a byproduct of child-care techniques. Table 8.4. Distribution of Cranial Modification  

 

Context

N

n

TPS-Sacrifices

Pacatnamú Farfan2

Punta Lobos

3

%

n

%

89

28

31.5

61

68.5

36

9

25.0

27

75.0

Adults

53

19

35.8

34

64.2

 

 

 

 

 

Túcume Burials

1

Unmodified

Subadults  

 

Modified

108

48

44.4

60

55.6

Subadults

35

16

45.7

19

54.3

Adults

73

32

43.8

41

56.2

 

 

 

 

 

96

38

40.0

58

60.0

41

39

95.0

2

5.0

91 

75

82.4

16

17.5

1. Verano (1986b) 2. Toyne (2006)  3. Verano and Toyne (2011)

A

b

23 6   Toyne

c

d

8.10. Variation in artificially modified cranial shapes at the Temple and within the Túcume cemetery: (A) moderately modified; (B) unmodified; (C) moderately modified; and (D) unmodified. Photos by J. M. Toyne.

Discussion Health, Identity, and Social Status Overall, the frequencies of paleopathological conditions that ultimately stem from social environments and experience suggest that among the sacrifice victims there was a degree of group homogeneity; however, these individuals were simultaneously dissimilar to cemetery burials at Túcume. While the mortuary context provides few clues as to the cultural origins of these sacrifices, the patterns in non-specific stress indicators and traumatic injuries suggest these individuals lived lives of greater biocultural stress than other individuals interred at Túcume. Beyond Túcume, these results can be compared to other contemporaneous Late Intermediate Period samples from the north coast (table 8.2).2 In the Jequetepeque valley, Farfán includes higher social status burials based on the complexity of grave offerings (Carol Mackey,

Life Before Death  237

personal communication; Toyne 2006), while Pacatnamú is a sample of non-elite individuals (Verano 1997a). Punta Lobos (Huarmey valley) is a large sample of ritually killed individuals that possesses a similar age and sex distribution as those killed at Túcume (Verano and Toyne 2011; Verano and Phillips, chapter 9). Demographic structures show the TPS sacrifices were clearly a select group restricted to children and young adult males—and it is notable that the other burials from Túcume do not include a substantial contribution from these demographic categories. This is not to say that the TPS sacrifices were directly selected from the local population, but this important observation may suggest either that young adult males at Túcume were interred elsewhere, or that they did not demonstrate the same risk of mortality as others living at the site (Bishop and Knüsel 2005). When compared to other burials at Túcume as a group, the TPS sacrifices demonstrate higher frequencies of nonspecific indicators of biocultural stress. Overall, the sacrificed juveniles demonstrated a higher proportion of active orbital lesions. This is consistent with the argument of Wood et al. (1992) that the cemetery individuals were less stressed or had perished before developing osseous markers of stress. However, there is a slight age bias in the non-sacrifice group toward younger individuals (but no infants). Porotic hyperostosis, in contrast, does not share the same pattern, and sacrifices as well as traditional burials demonstrated similar distributions with healed lesions, with a high prevalence among adult males. This marker would suggest that the victims were no less healthy than the rest of Túcume. Furthermore, porotic hyperostosis was generally elevated in both adult groups, which may represent different types of survived childhood stress, whereas cribra orbitalia was less prevalent and more likely linked to just childhood stresses. Overall, the high frequency of linear enamel hypoplasias across the entire sample (79.3%) stands in contrast to the lower prevalence of similar defects in inferred high-status burials at Farfán (Toyne 2006) and other Moche and Sicán elites (Klaus et al., in press; Shimada et al. 2004; Verano 1997b). Almost all skeletal indicators suggest the Túcume sacrifice group was under higher levels of chronic biological stress during their earlier years of life, and such elevated morbidity is consistent with lower social status. Interestingly, there is no difference in estimated statures across any of the samples considered, although the adult men from Farfán demonstrated an overall narrower and taller range. This was an unexpected

23 8   Toyne

pattern, since adult height could be reduced with the types of observed stresses that marked the teeth and bones. The pattern of lifestyle and cultural modification also led to different interpretations of the social experience and possible origins of the sacrificed individuals. While the TPS victims demonstrated almost twice as many healed postcranial traumatic injuries, no significant difference in craniofacial trauma exists to suggest that interpersonal violence was more prevalent in one group or another. Elevated prevalence of postcranial accidental trauma suggests not only a more physically active or hazardous lifestyle among adults but also that juveniles survived a greater number and variety of physical injuries. Finally, cranial modification does not appear to indicate any difference related to social identity among the groups. Moreover, cranial modification did not seem to have a particular social meaning for distinguishing social entities or statuses, or that the sacrifice group came from a similar group as observed in the Túcume cemetery. While the distribution of cranial modification was similar to those from Pacatnamú, the majority of Farfan’s Late Intermediate Period burials of higher inferred social status were modified (fronto-occipital style). As such, some particular variation in expression may exist and requires further exploration (table 8.2). Cranial deformation patterns at Punta Lobos are perhaps most instructive for Túcume. In that mass execution, clear variability in cranial modification style was found. This suggests that the Chimú executed a diversity of people on that Huarmey valley beach whose variety of cranial shapes indeed reflect diverse geographic or social origins (Verano and Toyne 2011). While this analysis strongly suggests that the TPS sacrifices represent a demographically select group, the evidence of childhood stress and patterns of antemortem trauma do distinguish them from the cemetery sample from Túcume (Toyne 2008). But they also differ in several ways from the burials of inferred higher social status from Farfán. Research in other sacrificial contexts similarly suggests that sacrificed individuals experienced greater overall biological stresses throughout their subadult life courses and that they were likely of lower social standing (Klaus et al. 2010; Klaus et al., chapter 7). Given their circumstances, victims may have had little choice in their selection, through either social or religious coercion, in the matter of becoming a sacrificial offering. Moreover, ritual deaths may have also served to establish and maintain positions of power by presiding elites.

Life Before Death  23 9

Models of Andean Human Sacrifice The osteological data from the sacrificial victims at Túcume allow us to model their life experiences and then compare them to sacrificial practices across the Andes, including earlier Moche sacrifices and later ethnohistoric documentation of Inka practices. While these comparisons are cross-cultural and based on secondary sources of information, including iconographic representations and ethnohistoric documents, bioarchaeological expectations can independently evaluate possible similarities and differences to known practices of ritual killing. The iconographic, archaeological, and biological evidence each strongly suggest sacrifice victims of the Moche culture (AD 100–800), especially at Huaca de la Luna’s Plaza 3A, were warriors who had experienced physical combat during their lives prior to sacrifice. The accumulated antemortem fractures exhibited in these men reflect a history of violent encounters (Verano 2001b:118). Most Moche victims were healthy robust young adult males, 15 to 39 years old, with pronounced muscle attachment areas (Verano 2000). The Moche victim profiles are clearly dissimilar to those at Túcume, especially in the high prevalence of stress markers and in the high representation of children and a few older individuals who all lacked consistent patterns of combat-related antemortem trauma. Other comparative models emerge from ethnohistoric descriptions of Inka human sacrifice during the Late Horizon Period in the southern Andean highlands (ca. AD 1430–1532), summarized in Besom (2009) and based primarily on Arriaga (1968 [1621]), Betanzos (1996 [1557]), Cieza de Leon (1963 [1538]), and Cobo (1990 [1653]). Several types of human sacrifice were identified, with archaeological and biological correlates that can be extracted from these accounts: • Retainer sacrifice produced bodies (often young women) to accompany deceased rulers in death—essentially as mortuary offerings, such as earlier Moche tombs at Sipán (Alva and Donnan 1993; Verano et al. 2001) and Sicán (Shimada et al. 2004). The TPS, however, was neither directly associated with a high-status tomb, nor were females present. Also, the TPS did not represent a mass immolation but rather the accumulation of offerings over several centuries. Furthermore, the patterns of sharp force trauma seen at the TPS, including throat-slitting, decapitation, and

2 4 0  Toyne

chest-opening, is inconsistent with any known form of retainer burial (Bentley and Klaus, chapter 10; Tomasto el al., chapter 11). • Capac hucha, or child immolations, involved physically “perfect” young children (boys and girls aged 10 into their teens) who were sacrificed, often in pairs, in elaborate ceremonies at high-altitude shrines. Burials included high-quality grave-good offerings to the sun and local mountain deities (apus) (Ceruti 2004; Reinhard 2005). While many of these ritual child killings occurred during the Inka occupation of the site, the TPS sacrifices appear inconsistent with Inka practices of capac hucha (Andrushko et al. 2011; Reinhard and Ceruti 2010). This assemblage included a large number of adult men, adults and children were treated similarly, and grave goods were absent.3 Many TPS individuals also demonstrated physical deformities that would have made them less than physically perfect. • The TPS killings could represent captured warriors executed by the Inka—usually decapitated—and buried without honor (Betanzos 1996 [1557]; Cieza de Leon 1959 [1553]). However, at the TPS, bodies were individually wrapped and buried in a coherent and careful manner in front of the temple. Patterns in cut marks reflect symbolic actions of throat-cutting and chest-opening to suggest intensely ritualized deaths, rather than a punitive killing where disgracing or destroying an enemy was the motivation (Duncan 2005). The presence of so many children (putative noncombatants) also makes warrior execution an unsupportable interpretive option. • Substitute sacrifices were killed to take the place of others who were usually sick in some way. The purpose was to exchange the death of a child in order for a parent to live, or for someone to take the place of another unwilling sacrifice (Cobo 1990 [1653]). This form of sacrifice leaves no physical or material correlates and is hard to substantiate. Still, it cannot be ruled out at Túcume considering as well the context and its llama offerings. This would suggest that sacrifices were selected from a vulnerable group in order to serve as an alternative sacrifice offered by individuals in positions of power. As suggested by taphonomic indicators and the number of individuals, they were regular propitiatory rituals (Toyne 2011); thus, avoiding selection for ritual death may have been important. • Runa (able-bodied male citizens) were also sacrificed in select ceremonies during the Inka Empire (Cieza de Leon 1959 [1553]; Las Casas 1967 [1550]; Xerez 1872 [1543]). The motivation and selection process in the ethnohistoric literature portrayed runa killings as oracular offerings and

Life Before Death  2 4 1

voluntary sacrifices. These sacrifices were supposedly dispatched by various methods including strangulation, decapitation, and heart removal. The TPS victims may present an earlier version of this type of sacrificial practice later adopted by the Inka. However, children were present and treated in a similar fashion as the adults at the TPS, suggesting, perhaps, that constructs of social value were not defined based on age and that younger individuals were able-bodied enough and thus sacrificed.

Ultimately, the killings at Túcume may have embodied a mix of sacrificial practices, such as runa and substitute sacrifices, and the temple patio was a specific disposal area for these distinct ritual offerings. In the end, however, secondary sources are Inka-specific texts and provide only limited details for a number of different types of sacrificial practices, including victim profile and manner of death. None of this evidence fits precisely the pattern presented at the TPS, except runa sacrifice. Ethnohistoric analogies also fail to substantively identify aspects of the victims’ social status beyond only that ritual death could be either punitive or a specific honor that could shift the victim’s social standing in either direction (Andrushko et al. 2011; Wilson et al. 1999). All of these lines of evidence converge on an understanding that late pre-Hispanic sacrifice on the north coast of Peru involved wholly unique and regionally distinctive practices. Many of these examples on the north coast present unique features that bespeak a remarkable variability in performance and practice, including victim selection and disposal. It is possible to reject a number of potential hypotheses about the origins of the TPS victims, including that they were captured warriors coming from great distances, that they were exemplars of physical perfection according to Inka standards for retainer burials, or that they were capac hucha child sacrifices. There was no clear evidence of physical or violent coercion (i.e., bindings) used to subdue these individuals. This research suggests that the victims all likely came from a local group of lower socioeconomic status and that they understood and possibly accepted the ritual process. It might also be considered that victims were in an altered state due to the consumption of alcoholic or narcotic substances that were provided by the sacrificers. The former is possible, and this scenario is further supported as the seeds of a known narcotic, anticoagulant, and hallucinogen, Nectandra sp., were recovered from within both the temple and nearby offering pits on the patio (Narváez 1995b; also Klaus et al. 2010; Montoya 1996; 2004). Ethnohistoric sources further

2 4 2   Toyne

provide descriptions supporting the notion that victims were drugged with such substances (Toyne 2015a; Xérez 1872 [1543]).

Conclusion This chapter examined how multiple skeletal pathological conditions can create a health profile of individuals selected for sacrificial deaths and assist in the exploration of multiple aspects of their social identity. The perimortem treatment of individuals from this bioarchaeological investigation suggests both a commonality in ritual practice and the mode for selecting individual offerings. Demography and pathology data at the Templo de la Piedra Sagrada demonstrate a selection pattern of victims that was strongly gender-based toward males and less influenced by age. Victims exhibited similar types and frequency of physiological stresses, but at slightly different rates from others interred at Túcume. The frequency of traumatic injuries resulting in visually identifiable disabilities may have been an important discriminating factor, suggesting a more difficult life, but this evidence alone does not offer a clear definition of their origins. As well, patterns of cultural modification of crania do not differ significantly from other locally similar cultural tradition, if not social standing. These pathological data suggest a profile of lived experience and biological health that was distinct from the patterns exhibited among the cemetery sample of inferred higher social status burials at Túcume. Overall, it is possible to conclude that the TPS sacrifices represented a group that experienced both social and environmental factors that negatively affected their health. It remains a clear possibility that sacrifice and offering of males and children of lower social status perhaps functioned as a means of maintaining social disparity and legitimizing broader social inequality in ways that still remain poorly understood. The study of these skeletal remains is just beginning. This analysis represents but the first step in the broader investigation of these sacrificed individuals. Additional lines of evidence for future studies include biodistance analysis of phenotypic expressions of cranial and dental traits (e.g., Nystrom 2006; Sutter and Cortez 2005) and isotopic analyses of oxygen and strontium signatures to examine possible geographic origins (Hewitt 2013). Carbon and nitrogen signatures will also aid in further reconstructions of dietary variability (e.g., Spence et al. 2004; White et al. 2009). Ancient DNA analysis—examining the patterns of

Life Before Death  2 4 3

kinship at the genetic level—may provide additional fine-grained perspectives as to identity and possible relationships to local and nearby populations (Lewis et al. 2004; Shimada et al. 2005b). Indeed, complexities and ambiguities in the archaeology of ancient identity may always remain, but a multidisciplinary bioarchaeological approach holds the best and most detailed perspectives for understanding life before death in sacrificial assemblages. Acknowledgments Funding was provided by the Social Science and Humanities Research Council of Canada (Grant No. 752–2004-0603) and BBC Television, U.K. Hearty thanks to Bernarda Delgado and Alfredo Narváez for archaeological support, and also to Natalia Guzmán, Sarah Baitzel, Elvis Mondragón, Victor Curay, and Oswaldo Chozo for field and lab assistance. Carol Mackey provided access and permission to use materials from Farfán, and John Verano was the principal investigator of the Punta Lobos collection. This chapter was strengthened by the comments of the anonymous reviewers and my colleague and coeditor, Haagen Klaus.

Notes

1. Subsequent pathological analyses are based on frequency in specific skeletal elements, such as cranial or dental remains, and not based on the MNI, as numerous individuals were incomplete. 2. I have personally collected the data for these samples. In the case of Pacatnamú, these data were collected and published by my dissertation adviser, Dr. John Verano, who supervised my research. 3. There were three small Inka figurines of silver and Spondylus included as offerings associated with the Temple of the Sacred Stone, similar in all respects to those found with the high-altitude offerings of capac hucha children (Narváez 1995b). However, there was no direct association with any specific sacrificial offerings. The mortuary and perimortem manipulation was quite distinctive at Túcume from known capac hucha (Reinhard 2005).

2 4 4   Verano and Phillips

chapter nine

The Killing of Captives on the North Coast of Peru in Pre-Hispanic Times Iconographic and Bioarchaeological Evidence John W. Verano Department of Anthropology, Tulane University

Sara S. Phillips Hopkinsville Community College

Iconographic and archaeological evidence provides clues to interpreting warfare and other forms of violence in ancient cultures. Artistic depictions of armed warriors and bound figures, which we will refer to as captives, first appear on the north coast of Peru during the Initial Period (ca. 1800–900 BC). Perhaps not coincidentally, the earliest skeletal evidence of interpersonal violence in this region is found during this period as well (figure 9.1). Combat and the taking of captives become dominant artistic themes in the Moche culture (ca. AD 100–850) (Donnan and McClelland 1999; Donnan 2010; Uceda, Tufinio, and Mujica 2011). In the later pre-Hispanic cultures, depictions of captives are less frequent, but examples are known in Chimú art (Jackson 2004; Lapiner 1976; Uceda 1999). Archaeological evidence of the systematic killing of captives also first appears in Moche contexts (beginning ca. AD 200) and continues until the collapse of Moche society (ca. AD 850). Skeletal remains of executed captives (discussed in this chapter) also have been found at several north coast sites dating to the Late Intermediate Period (ca. AD 1100–1350), indicating that such practices continued until late pre-Hispanic times. In this chapter, we examine the killing of captives within the larger context of warfare and state expansion in northern coastal Peru. We discuss four examples of captive sacrifice in order to identify the

The Killing of Captives, North Coast of Peru  2 45

9.1. Skull of an adult male with multiple healed cranial wounds. From a Cupisnique burial (Paredones Tomb 2, Burial 3B) at the site of El Brujo in the Chicama Valley (Mujíca et al. 2007).

unifying characteristics that distinguish it from other forms of sanctioned killing, such as retainer sacrifice or the offering of human lives for religious purposes (see Verano 1995). Instead, we see the killing of captives occurring in a larger political context of territorial expansion and control, with the most important diagnostic elements being exclusively male victims, evidence for violent death, the presence of physical restraints, and the denial of proper burial (see discussion below). In seeking to understand the motivations behind the killing of captives, we will use archaeological data to establish their identity as captives. We will then use biological profiles and the individuals’ antemortem trauma pattern (the presence and frequency of healed fractures) to establish their likely identity as either captured enemy soldiers or non-combatants (Phillips 2009). Though we consider each of the presented examples to be cases of captive sacrifice, the profiles of the victims suggest the motivation of the practice differed over time and among north coast cultures.

Identifying Captives in the Iconographic and Archaeological Records In many Andean artistic traditions, captives are recognizable as bound male figures (figures 9.3, 9.8, 9.9). They are typically shown stripped

2 4 6   Verano and Phillips

not only of clothing but of all adornments and signifiers of rank such as headdresses, nose ornaments, and ear spools. In the case of ear spools, their absence is made obvious by the depiction of visibly pierced (but empty) ear lobes (Donnan 2004). Captives may be depicted as single individuals or in larger groupings showing their subjugation, presentation to high-status figures, and in processions led by ropes around their necks (Donnan 2010). We are not aware of any depictions of women or children as combatants or captives, although women do appear in Moche art as active participants in the sacrifice of male captives (Alva and Donnan 1993; Donnan and Castillo 1994). Direct archaeological evidence of the killing of captives includes skeletal or mummified remains with ropes around wrists, ankles, or the neck (figures 9.6, 9.8, 9.9); evidence of violent death includes fractured skulls, slashed throats, penetrating wounds, or the cutting open of the chest (figures 9.7, 9.12) (Verano 1986a, 1995, 2001, 2008). Burial context is important, if not key, as well. Captives have not been found in tombs, but they have been discovered buried in architecture, deposited on the surface of plazas or in the bottom of trenches, and buried in isolated locations away from settlements and ceremonial precincts (Bourget 2001b; Verano and Toyne 2011). Typically, the position and orientation of the bodies of sacrificed captives are highly variable (figures 9.4, 9.5, 9.10) and do not conform to normative burial practices. Identities of Captives In Andean iconography, captors and their captives may in some cases be distinguished from one another by differences in clothing or weapons (Lau 2004). Unfortunately, unlike the case of other complex New World societies such as the Aztecs or Maya, where important captives were identified by inscriptions or painted images giving their name and (in some cases) date of capture (e.g., Boone 2000: fig. 3; Tiesler and Cucina 2007), no writing or pictographic system capable of recording such information existed in pre-Hispanic South America. Thus, epigraphy cannot be used to contextualize Andean depictions of prisoner capture and sacrifice (Verano 2001b). The skeletal remains of sacrificed captives may reveal some information about their identities, however. Morphological or metric features of the skull and dentition, as well as stable isotopic analysis of bones and teeth and ancient DNA, can be used to test whether captives can be distinguished in some way from their

The Killing of Captives, North Coast of Peru  247

presumed captors and, potentially, provide some indication of their geographic origins, as will be further explored below.

Bioarchaeological Evidence of Conflict in Northern Coastal Peru Given that the capture and killing of captives is an outgrowth of interpersonal conflicts—whether small-scale raiding or organized warfare—we begin with a chronological review of the archaeological evidence of warfare and the taking of captives in northern Peru. The Initial Period (ca. 1800–900 BC) Although monumental architecture and the foundations of early complex societies in highland and coastal Peru have their origins in the Late Preceramic Period, it is not until the Initial Period that there is a marked increase in the construction of monumental ceremonial centers in northern coastal Peru (Burger 1992a; Pozorski and Pozorski 1992, 2002). Advances in metallurgy and ceramics at this time also produced the first great art style of northern coastal Peru, known as Cupisnique (ca. 1000–500 BC) (Larco Hoyle 1941). The Initial Period is marked by settlement pattern shifts and the development of irrigation agriculture, accompanied by a substantial increase in population in many northern coastal valleys. With such dramatic changes in social formations and settlement patterns, competition and conflict over resources would be expected and might be reflected in skeletal remains from Initial Period burials (figure 9.1). Unfortunately, looting and destruction of sites have resulted in loss of contextual information from thousands of Cupisnique burials that might have provided such evidence. Only a very small number of graves from this period have been excavated scientifically, and fewer still have detailed descriptions of the skeletal remains encountered in them (Alva 1986; Elera 1998; Mujica et al. 2007; Ravines et al. 1982). Indirect evidence of conflict is present, however, in the form of artistic depictions of warriors and mutilated victims at the site of Cerro Sechín in the Casma valley (figure 9.2a, figure 9.2b; Fuchs 1997), as well as modeled ceramics of bound captives in the Cupisnique (Tembladera) style of the Jequetepeque valley (figure 9.3; Alva 1986; Donnan 1992; Lapiner 1976). Unfortunately, the lack of large skeletal samples from this early period greatly limits our ability to document violence associated with conflict or ritual practices.

2 4 8   Verano and Phillips

A

b 9.2. Carved stone reliefs at Cerro Sechín, Casma valley: (A) image of a warrior holding a club and (B) carved image of two columns of severed heads. Photos courtesy of Julio Vizcarra.

9.3. Cupisnique (Tembladera) ceramic vessel in the form of a bound captive. Private collection, Peru. Photo courtesy of Alana Cordy-Collins.

The Killing of Captives, North Coast of Peru  2 49

The Early Intermediate Period and Middle Horizon Of all pre-Hispanic societies of Andean South America, the Moche (ca. AD 100–850) were the most prolific illustrators of warfare and the taking and killing of captives. Moche art is highly representational, known for its realistic depictions of humans, plants, and animals in threedimensional modeling and fineline slip painting on ceramics, as well as on metalwork and textiles. One of the most common themes in Moche art is what Christopher Donnan has termed the “Warrior Narrative,” a series of depictions that include warriors preparing for and engaged in combat, taking captives, and sacrificing them in elaborate ceremonies (Alva and Donnan 1993; Donnan 2010; Donnan and McClelland 1999). Weapons are shown prominently in this narrative, both in the hands of combatants and in the form of weapon bundles, which can stand alone as metaphors for combat and the capture of opponents. However, the question of who exactly these opponents are continues to be an issue for debate. While some of these combat scenes show confrontations between supernatural figures (presumably Moche deities or other supernatural antagonists) and are clearly metaphorical, the majority depicts combat between human warriors, albeit in highly formalized scenes (Alva 2001; Quilter 2002, 2008; Verano 2001b). A problematic issue for many years was that the tombs of the elaborately dressed figures shown presiding over the arraignment and killing of captives, as well as the remains of the victims themselves, were seemingly absent from the archaeological record. This absence of evidence led some to suggest that Moche warfare and the killing of captives was largely a mythical narrative and not a representation of actual practices (Alva and Donnan 1993). However, beginning in 1987 at the site of Sipán, a series of tombs was excavated in the Lambayeque and Jequetepeque river valleys of individuals dressed in the elaborate garb of those who presided over the sacrifice of captives in Moche art (Alva and Donnan 1993; Donnan and Castillo 1994). In 1995, the first discoveries of the remains of sacrificial victims were made at Huaca de la Luna in the Moche river valley (Bourget 1997, 1998b, 2001). Subsequent excavations uncovered additional deposits of victims at the Pyramids of Moche (Verano 2008a; Verano, et al. 2008). With these important discoveries, depictions of combat and the execution of prisoners in Moche art could finally be compared directly with skeletal evidence and the archaeological contexts in which the remains

2 5 0  Verano and Phillips

were found. Additional evidence of warfare comes from settlement pattern studies that document the construction of numerous fortified hilltop sites throughout the Moche domain, indicating that defense was a significant concern, at least during certain periods (Billman 1997; Dillehay 2001; Wilson 1988). The Pyramids of Moche Archaeological excavations at the Pyramids of Moche site have revealed the remains of more than 150 executed captives, with the largest number in two small courtyards (Plazas 3A and 3C) on the east side of Huaca de la Luna (Pyramid of the Moon) (figures 9.4, 9.5, 9.6). In both contexts, all appear to be skeletally healthy males between the ages of approximately 15 to 40 years. Some individuals have antemortem fractures of the arm, ribs, or shoulder blade that were in the early stages of healing at the time of death (Phillips 2009; Verano 1998, 2001a, 2001b, 2001c). One had a fracture of the occipital bone that showed initial healing as well. These injuries appear to mark wounds received in combat or following capture. They are consistent with Moche depictions of armed battle, where the primary objective seems to have been to disable and capture, rather than kill, one’s opponent (Alva and Donnan 1993). These healing

9.4. A photo taken early in the excavation of Plaza 3A showing some of the initial discoveries of sacrificed captives. Photo by John Verano.

The Killing of Captives, North Coast of Peru  2 51

9.5. Skeletons of sacrificial victims incorporated in the construction fill beneath the floor of Plaza 3C. Photo by John Verano.

9.6. Sacrificed bound captive in Plaza 3C found with fragments of rope encircling the wrists and neck. Photo by John Verano.

2 5 2   Verano and Phillips

9.7. Cervical vertebrae C2–5 from Plaza 3C Entierro 2. Multiple cut marks are visible across the anterior surface of the body of C3 and the transverse processes of C2–C4. Photo by John Verano.

fractures are also important in indicating that some time (at least several weeks, perhaps more) elapsed between the moment of capture and death (Verano 2001c). Rope fragments were found around wrists and ankles of some victims (figure 9.6), indicating that they had been physically restrained. One skeleton in Plaza 3C had the hands tied behind the back and joined to a rope encircling the neck. These bindings are consistent with depictions of captives in Moche art. Moche artistic depictions of prisoner execution show victims having their throats slit and their blood collected for presentation to an elaborately dressed figure (Alva and Donnan 1993; Castillo and Holmquist 2000). Skeletal evidence from Plazas 3A and 3C is consistent with the art: the most common perimortem injuries seen in these skeletons are cut marks across the bodies and transverse processes of the cervical vertebrae (figure 9.7). The location of the cut marks, limited in the majority of cases to the anterior surfaces of the vertebrae, indicates that the objective was to slit the throat (Hamilton, chapter 2) and not to decapitate the victim, although a few individuals were decapitated and had their skulls modified as trophies (Verano et al. 1999).

The Killing of Captives, North Coast of Peru  2 53

Fragments of ceramic vessels in the form of prisoner figures were found associated with the sacrificial victims in Plazas 3A and 3C. In the case of Plaza 3A, these are large unfired clay effigy figures in the form of seated nude males with ropes around their necks (Bourget 2001b; Donnan 2004). Plaza 3C contained fragments of smaller fired vessels in the form of seated prisoners with ropes around their necks and hands bound behind their backs (figure 9.8). All vessels from Plaza 3C were fragmentary, but two were complete enough to allow reconstruction of their original form (Armas 2008). Given their direct association with the skeletal remains of sacrificial victims, the ceramic vessels found in Plazas 3A and 3C played some role in the sacrificial rituals, and their intentional breakage seems to have been an integral part of the ceremony.

9.8. Drawing of a prisoner vessel reconstructed from multiple fragments. Plaza 3C, Huaca de la Luna. Drawing by José Armas.

2 5 4   Verano and Phillips

Source of Captives and Motivation for Killing The source of Moche captives remains a subject of debate. While mtDNA studies of a small sample of Plaza 3A victims suggest that they are related to individuals buried at the Moche site (Shimada et al. 2008; Shimada et al. 2005), dental morphological analyses of a select range of nonmetric traits reach a different conclusion, classifying the sacrificial victims as outliers distinct from the local valley population (Sutter and Cortez 2005; Sutter and Verano 2007). As has been argued elsewhere (Verano 2001b, 2008), sacrificial victims at Huaca de la Luna probably represent captives from battles fought between different Moche groups and possibly conflicts with non-Moche polities. An alternative view— that Moche warfare was not warfare at all but ritual combat between the elites of the Moche site (Topic and Topic 2009)—is not, in our opinion, supported by the bioarchaeological evidence. The specific motivation for killing of captives by the Moche also remains a subject of significant discussion among Moche specialists. Steve Bourget hypothesizes that prisoners were sacrificed in Plaza 3A in times of crisis associated with torrential El Niño rains, since some of the victims appear to have been killed during periods of heavy rainfall (Bourget 1997, 1998b). Excavations in Plaza 3C revealed evidence of occasional rainfall that deposited thin layers of silt over the plaza, although none of the skeletons were directly associated with these layers. Thus Plaza 3C does not provide evidence to support a direct relationship between El Niño rains and human sacrifice. Given the three temporally distinct deposits of victims at Huaca de la Luna (Plaza 3C subfloor; Plaza 3C above floor; Plaza 3A), it is now evident that these events occurred over a period of centuries. The killing of captives was clearly a long-standing cultural tradition that played an important role in ritual activities at Huaca de la Luna and was not exclusively a response to episodic environmental crises such as the El Niño phenomenon. While sacrifices in Plaza 3A may have been made for different reasons and on a different scale than those in Plaza 3C (Uceda and Tufinio 2003), the type of victims chosen and the way in which they were killed are the same. Patterns in Antemortem Trauma Along with the similarities in perimortem trauma and postmortem treatment, the captives from Plaza 3A and Plaza 3C are also similar in terms of their antemortem trauma pattern. The individuals from both plazas show evidence of numerous healed fractures to the frontal and

The Killing of Captives, North Coast of Peru  2 55

left parietal bones, which are suggestive of a history of armed combat in the years before their deaths (Phillips 2009). Plaza 3A had an antemortem frontal bone fracture prevalence of 18.6%; Plaza 3C’s was 13.3%. The prevalence for left parietal fractures was lower, at less than 7% for both samples. Fractures of the frontal and left parietal suggest face-toface combat with an individual wielding a blunt weapon, such as a club, in the right hand. Well-healed fractures were also found on ribs and several bones of the upper and lower limbs in both samples. Individuals from the two contexts had a prevalence of left distal ulnar fractures comparable to other bioarchaeological samples linked to warfare, around 5.5% for Plaza 3A and 4.0% for Plaza 3C, including healing fractures (Phillips 2009). Though the interpretation of ulnar fractures as always indicative of violence has been challenged (Lovell 1997), all of the ulnar fractures observed in the Huaca de la Luna samples are located in the distal portion of the left ulna and are generally transverse in nature, which argues for a diagnosis of a true parry fracture (see Judd 2008 for a detailed review of parry fractures). Moche iconography sometimes shows warriors carrying shields on their left arms, suggesting they indeed would use the left arm to parry blows. In general, the older individuals from the Pyramid of the Moon show a greater number of fractures than the younger, indicating they had experienced more violent episodes over the course of their lifetime. While the overall pattern of antemortem trauma is similar between the two samples, there are some differences between the Plaza 3A and Plaza 3C individuals. The overall number and frequency of antemortem fractures is greater in the Plaza 3A samples, including those to the anterior region of the skull. Additionally, fractures are found in a wider variety of places, including facial fractures and a greater number of fractured ribs (Phillips 2009). The Plaza 3A group appears to have experienced a greater level of violence through the course of their lives than was true for the individuals in Plaza 3C. Importantly, comparative samples of Moche burials from the Pyramid of the Moon and other Moche sites show much lower frequencies of bone fracture than those seen in the Plaza 3A and 3C groups (Phillips 2009). Late Intermediate Period (ca. AD 1000–1470) With the collapse of Moche society around AD 850 came major social reorganization and the emergence of new cultures, marked by changes in burial practices, architecture, settlement pattern, and art styles

2 5 6   Verano and Phillips

9.9. Chimú wooden prisoner figure from Huaca Tacaynamo, Moche valley, with original cord bindings preserved. Height: 51cm. Courtesy of the Archaeology Museum of the National University of Trujillo, Peru.

(Donnan and Mackey 1978; Moseley 2001). Following a transitional period that remains poorly understood, two major polities emerged: the Lambayeque (or Sicán) in the northern valleys, and the Chimú, centered in the Moche valley (Moore and Mackey 2008; Shimada 1990). Both the Lambayeque and Chimú polities extended their domains to encompass multiple coastal valleys, and by the middle fourteenth century a series of Chimú military campaigns culminated in the political domination of the entire north coast (Moseley and Cordy-Collins 1990; Rowe 1948). Neither would develop art styles with realistic portrayals of human figures like those of the Moche; humans generally appear as generic and stylized figures, and there are no depictions of armed combat (Donnan 1992). The only apparent continuity is in representations of captives, of which some Chimú examples are known (figure 9.9; Jackson 2004; Uceda 1999; Verano 1986a). Archaeological evidence of the Chimú military conquests of north coast valleys is largely indirect, in the form of intrusive administrative centers built in valleys to the north and south of the Chimú capital of Chan Chan (Mackey and Klymyshyn 1990;

The Killing of Captives, North Coast of Peru  2 57

Moore and Mackey 2008). Physical remains of battle victims have not been found, although the extensive looting of archaeological cemeteries has resulted in a relatively small sample of documented human remains from this period. Two exceptions are a mass burial in a defensive trench at the site of Pacatnamú in the Jequetepeque valley, and a large-scale execution of captives at Punta Lobos in the Huarmey valley. Pacatnamú In 1984, fourteen human skeletons were found at the bottom of a 3 meter–deep defensive trench at the archaeological site of Pacatnamú in the lower Jequetepeque valley (Verano 1986a). They were found at the entrance to a major architectural complex known as Huaca 1 (figure 9.10) that was built during the later phase of occupation of the site, circa AD 1100–1400 (Donnan 1986). The fourteen individuals were all adolescent and young adult males, ranging in age from approximately 15 to 35 years (Verano 1986a). They were found in three superimposed groups, each group separated from one another by a layer of sand and rubble. Surface weathering on some of the bones, as well as abundant remains of scavenging insects, indicate that the bodies were not promptly buried but rather decomposed on the surface (Faulkner 1986). Rope fragments were found around the ankles of one individual (figure 9.11) and around the wrist of another, indicating that the victims were bound or hobbled. No traces of clothing other than a possible loincloth fragment were found with the skeletons. Laboratory study revealed evidence of multiple injuries, including stab wounds, cut marks, skull and long bone fractures, and forced dismemberment. Two individuals had cut marks indicating that their throats were slit, two were decapitated, and five appear to have had their chests cut open (Verano 1986a). In addition, four individuals were missing their left radius. Cut marks and fractures on bones that articulated with the missing radii indicate that the bone was intentionally removed from the victims. The radii may have been collected as trophies to be modified as flutes or other objects; there are descriptions of similar practices in Inka times (Ogburn 2007; Verano 1995). The identity of the victims in the mass burial is unknown, although the ropes around their ankles and their age and sex distribution (nearly identical to those of the Huaca de la Luna victims) suggest they were captive warriors. Radiocarbon dates and their context (associated with the Late Intermediate Period occupation) indicate a date for the deposit

2 5 8   Verano and Phillips

9.10. Oblique view of the uppermost group of the Pacatnamú mass burial victims in the bottom of a 3m-deep defensive trench. Photo by John Verano.

9.11. Preserved rope around the ankles of a victim from the Pacatnamú mass burial, Jequetepeque valley. Photo by John Verano.

The Killing of Captives, North Coast of Peru  2 59

of around AD 1100–1200 (Verano and DeNiro 1993), contemporaneous with the estimated date of conquest of the Jequetepeque valley by the Chimú (Moore and Mackey 2008). Biometric comparisons of the victims’ crania with samples from Pacatnamú and nearby coastal and highland sites were not conclusive in identifying them either as members of the local population or as outsiders. However, carbon and nitrogen isotopic analysis of their bone collagen indicated that half (those in the lower two groups of the mass burial) had a non-marine diet distinct from that of the local Pacatnamú burial population, suggesting that they were not locals (Verano and DeNiro 1993). As with the captives from the Pyramids of Moche, there are some suggestions that at least a few of those recovered from the Pacatnamú mass burial were experienced combatants. Two individuals had healed fractures of the ribs, and another displayed a large healed depressed cranial fracture. While this represents a lower absolute number of fractures than were observed in Plaza 3A and Plaza 3C, the Pacatnamú sample size was also much smaller. The prevalence of antemortem cranial fractures, at 7%, is higher than the 2.9% prevalence found in a cemetery sample representing lower- to middle-class inhabitants of Pacatnamú, whose general fracture patterns were consistent with occupational trauma (Phillips 2009; Verano 1997a). Punta Lobos To date, the largest sample of what appears to be a mass execution of bound captives found anywhere in Peru comes from a Late Intermediate Period site at Punta Lobos, located in the Huarmey river valley (Verano and Toyne 2011). The Punta Lobos site was discovered in 1998 on a hill overlooking the Pacific Ocean by archaeologist Hector Walde (1998). His excavations recovered the remains of nearly 200 individuals buried at a shallow depth. Previous clandestine excavations had disturbed a portion of the site, but more than 100 bodies were found in situ. Their wrists and ankles were bound with rope or cloth, and cloth blindfolds were found still in place on many of the individuals. The throats of all victims had been slashed repeatedly, as indicated by multiple cut marks across the lower cervical and upper thoracic vertebrae, and on the clavicles and first ribs (figure 9.12). The great majority of victims were found lying facedown in the sand, suggesting that they had been forced into this position and then killed by someone straddling the victim from behind. Other than a few fragments of Spondylus shell, no offerings were found directly associated with the bodies. A small pit found on an

2 6 0  Verano and Phillips

9.12. Cut marks on the body of the first thoracic vertebra of a victim from Punta Lobos, Huarmey valley. Photo by John Verano.

adjacent hillside contained simple offerings, including ceramics of a local style, a fishing net, and food, possibly a gift left for the victims by relatives. The Punta Lobos mass execution is unusual compared to other sacrificial deposits described in this chapter for its location on an isolated hillside with no associated architecture. It is also unusual in that although most victims are young males, children as young as seven years and old men are also present (Verano and Toyne 2011). This is a different demographic profile from that seen at Pacatnamú and at the Pyramid of the Moon, where all victims were adolescent or young adult males—an age range appropriate for captured warriors. The Punta Lobos victims also have a very different traumatic profile than that of the other groups of executed captives discussed above. Apart from cut marks, the only perimortem trauma found in the Punta Lobos remains was a single fractured right first rib. Several cut marks were present on this rib, suggesting the fracture resulted from blunt force trauma associated with the cutting of the throat (Phillips and Verano 2005). There were a number of fully healed antemortem fractures observed, including a few to the cranial vault and face, but the percentage of individuals with a healed cranial fracture was significantly lower than in the

The Killing of Captives, North Coast of Peru  2 6 1

sacrifices from the Pyramids of the Moon; the prevalence of antemortem fractures was only 3.4% for both frontal and left parietal fractures at Punta Lobos (Phillips 2009). There was also little overlap in the location of postcranial fractures between the two samples. For example, the one fracture observed in a left ulna at Punta Lobos was in the proximal portion of the bone, indicating it was not a parry fracture. However, there were a large number of rib fractures observed in the Punta Lobos victims, particularly to the sternal portions of left ribs. Interestingly, the rate of fracture for this location was higher in the Punta Lobos remains (between 1% and 2%) than in the Plaza 3A remains (0.5%); there were no sternal fractures of left ribs observed in Plaza 3C (Phillips 2009). Etiology of rib fractures and interpretation of their prevalence in bioarchaeological samples are not well understood, in no small part because the ribs present particular challenges in terms of inventory and observation. However, Lovell (1997) suggests posterior rib fractures are likely to result from violent interactions, while rib fractures in other portions are likely to be accidental or occupationally related. At Punta Lobos, the concentration of fractures in the sternal third of ribs suggests they are part of a pattern of occupational trauma and that the occupation was not armed combat. The isolated location of Punta Lobos, on a hillside overlooking the Pacific Ocean and distant from any ceremonial architecture, suggests an event quite distinct from the executions of captive warriors at the major ceremonial centers of Huaca de la Luna or Pacatnamú. Punta Lobos appears to represent a mass execution of another sort, possibly a reprisal killing. Radiocarbon dates place the event at around AD 1250–1300 (2 sigma calibrated), which coincides with the estimated date of conquest of the Huarmey valley by the Chimú (Mackey and Klymyshyn 1990). The Punta Lobos victims thus may represent a Chimú response to local resistance (Phillips and Verano 2005; Verano and Toyne 2011). The Late Horizon (AD 1470–1533) Although ethnohistorical sources describe the military defeat of the Chimú by the Inka Empire around AD 1470, no direct archaeological evidence of this conquest has been found to date. Evidence of the relatively brief Inka domination of northern Peru is largely ephemeral, consisting of scattered Inka offerings and burials and the appearance of new ceramic styles incorporating Inka motifs and vessel forms (Morris

2 6 2   Verano and Phillips

and von Hagen 2011; also Hayashida 1998; Tschuaner 2001). While it is known that the Inka took captives following military victories, to our knowledge there are no specific accounts of these from the Chimú conquest. Although the Inka-period sacrifices discussed in other chapters (see Klaus et al., chapter 7 and Toyne, chapter 8) may point to some other form of victim selection criteria, there are no clear indications those victims were captives.

Discussion: Motivations for the Killing of Captives Otterbein’s (2000) important cross-cultural study of the killing of captives highlights some of the motivations for such behavior. Unfortunately, the archaeological cases described in this chapter do not have associated ethnohistoric accounts that might be used to explain the context of these events. There are such sources available for the Inka, however, which may serve as a regional model for comparison. Spanish Colonial period authors recorded various ethnohistoric accounts of the execution of captives in celebrations following Inka military conquests, as well as reprisal killings in response to acts of resistance or rebellion (Agro 1972; D’Altroy 2002; Rostworowski de Diez Canseco 1999; Rowe 1946). A variety of trophies made by the Inka from the skulls, long bones, teeth, and skin of slain enemies was described by chroniclers (Ogburn 2007). While captured enemies may have been killed as “offerings” to gods or ancestors in religious ceremonies, the execution and mutilation of captives clearly functioned as a powerful means to humiliate, pacify, and terrorize enemies. Such was the case of the Inka conquest of the Collas, whose leaders’ heads were severed and placed in a special building in Cuzco along with those of other conquered enemies (Sarmiento de Gamboa 1942). Retaliatory massacres also served to cement Inka conquests and discourage resistance (Rostworowski 1999:73–79). The treatment of the bodies of executed captives may provide some insight into how these victims were viewed by their captors. Careful burial with sumptuary goods would imply that the victim had been transformed into an offering of value, as is the case in dedicatory burials under ramps and audencias at the site of Chan Chan (Andrews 1974; Day 1982), or in the case of Inka child sacrifices known as capacocha (Ceruti 2003; Reinhard and Ceruti 2000). In contrast, desecration of the corpse by mutilation, exposure to scavengers, and denial of standard ritual burial conveyed a distinct message. The Pacatnamú mass burial is a good example of the

The Killing of Captives, North Coast of Peru  2 6 3

latter. Here the mutilated remains of executed captives were thrown into the bottom of a trench at the entrance to a ceremonial precinct and were left exposed to flies and other scavengers (Faulkner 1986; Rea 1986; Verano 1986a). The prominent display of the decomposing bodies of the victims (figure 9.10) and the denial of proper burial was clearly intentional, as was likely the case at Punta Lobos as well. The intentional exposure of bodies at all three sites is one indication that they represent executed captives. Bindings, either of cloth or rope, were found around the hands, feet, or necks of individuals at all of the sites, indicating their controlled, captive status. It should also be noted that, as far as can be established through osteological analysis and ancient DNA (aDNA) analysis of subadults at Punta Lobos (Scola 2004), all the captives described here are males. Those from the Pyramids of Moche and Pacatnamú shared similar age profiles. They were all able-bodied adolescent or young adult males. The wider age range of Punta Lobos, including males as young as 7 years and many greater than 50 years, suggests that this was an event fundamentally different from the killings at the Pyramid of the Moon and Pacatnamú. Indeed, the killing of captives, as a practice, seems to be similar at Huaca de la Luna and Pacatnamú, as these sites share many more features than either does with Punta Lobos. There are differences in the specifics of the postmortem treatment of remains between the two sites, but they share a proximity to monumental architecture, evidence for the taking of trophies in the form of body parts, a very limited age range for the victims, and antemortem trauma profiles indicative of a history of armed combat. This contrasts with the victims at Punta Lobos. The greater age range of victims, the distance from any form of settlement or ceremonial center, as well as the lack of postmortem handling of remains and antemortem trauma patterns not consistent with armed combat all make Punta Lobos stand out. However, the dates of Pacatnamú and Punta Lobos suggest that both events occurred during the expansion of the Chimú polity. Indeed, all of these examples of captive sacrifice are linked to expansionist polities. At the Huaca de la Luna and Pacatnamú, those killed were likely captured warriors brought back to a major ceremonial center for sacrifice. Osteological evidence indicates that at least some of them were experienced fighters. The varying levels of antemortem trauma indicate that the Plaza 3A victims had a greater history of combat experience than

2 6 4   Verano and Phillips

all the other samples—including Plaza 3C—perhaps because betweengroup conflicts increased as the Moche polity began to fall apart. In contrast to these groups, the Punta Lobos victims appear more like civilians than experienced soldiers, suggesting they were rounded up and killed by the Chimú in response to resistance or rebellion.

Conclusion The three contexts of captive execution described here—Huaca de la Luna, Pacatnamú, and Punta Lobos—provide convincing archaeological evidence that captives were taken and killed by the Moche and Chimú and that prisoner execution was a persistent and pervasive practice. The specific motivations for killing captives can only be hypothesized, but differences among the places in which captives were killed and the demographic profiles of the victims suggest distinct motivations for the practice. Archaeological evidence clearly indicates that the Chimú were an expansionist empire (Mackey and Klymyshyn 1990; Moore and Mackey 2008). The nature of Moche political organization continues to be a subject of debate, with models ranging from a confederation of valley chiefdoms to a centralized state ruled from the Pyramids of Moche (Quilter and Castillo 2010). Regardless of which model one accepts, however, the warrior narrative in Moche iconography, joined by archaeological evidence of prisoner capture, sacrifice, and trophy-taking, clearly document the importance of militarism in Moche society. The taking and sacrifice of captives probably served similar functions for the Moche and Chimú: as an affirmation of political power and as a potent message to those they conquered. And archaeological and osteological evidence demonstrates that at least some of the depictions of combat and the sacrifice of captives seen in Moche and Chimú art depict real events and not simply mythological narrative. With new discoveries come new complexities. A series of recent discoveries of human sacrifices in the Lambayeque river valley of northern coastal Peru has complicated what for many years seemed to be a clear dichotomy between violent execution of captives and respectful treatment and careful burial of sacrificed women and children. New discoveries from the sites of Túcume (Toyne 2011a, 2011b, Toyne, chapter 8), Cerro Cerrillos (Klaus et al. 2010), and Chotuna (Turner 2013; Klaus et al., chapter 7) are puzzling in that they exhibit diversity in age and sex (children and females at some, only males at others), as well as evidence at

The Killing of Captives, North Coast of Peru  2 6 5

two sites of the careful burial in shrouds of victims who have had their throats slit (and in some cases were decapitated) and chests cut open. This apparent contradiction between respectful treatment of the body (careful burial in a shroud) and a particularly violent death does not match the patterns previously defined for human sacrifice and the killing of captives on the north coast of Peru. These discoveries also raise challenging questions about the identity of the victims and their sacrificers, as well as questions about the nature and meaning of these sacrificial rituals. Clearly the final chapter on this topic has yet to be written. Acknowledgments We would like to thank the directors of the following archaeological projects who made excavation and analysis of skeletal samples possible: the Pacatnamú Project (Christopher Donnan and Guillermo Cock), the Huaca de la Luna Project (Santiago Uceda and Ricardo Morales), and the Punta Lobos/Huarmey Valley Project (Hector Walde). We also wish to thank Steve Bourget, who directed the Huaca de la Luna Plaza 3A excavations, and Moises Tufinio, who co-directed the Huaca de la Luna Plaza 3C excavations. Thanks also to a dedicated team of graduate and undergraduate students from Tulane University who participated in the excavation of Plaza 3C and analysis of material during the 2000 and 2001 field seasons:  Laurel Hamilton, Cathy Gaither, Lori Jahnke, Anne Titelbaum, Ginesse Listi, Stan Serafin, Mary Sawyer, and Teresa Gotay. We are also grateful to Florencia Bracamonte and Moises Rivero of the Universidad Nacional de Trujillo, Mellisa Lund of the Universidad Nacional Mayor de San Marcos, and Melissa Murphy of the University of Wyoming for their participation in excavation and laboratory analysis. Tulane anthropology students Heather Backo, Julia Drapkin, Helen Rich, and Kris Krowicki assisted with excavation and analysis of skeletons during the 1999 excavations of Plaza 3C under the direction of Moises Tufinio. Funding for the 2000–2001 Plaza 3C excavations was provided by grants 6784-0 and 7024-01 from the National Geographic Society’s Committee for Research and from the Huaca de la Luna Project. The Roger Thayer Stone Center for Latin American Studies at Tulane University kindly provided summer research grants for graduate student travel. Julia Drapkin and Teresa Gotay received travel funding from the Kenneth J. Opat Fund, Department of Anthropology, Tulane University. Research at Punta Lobos was made possible by grants from the Roger Thayer Stone Center and from Engel Brothers Media. Thanks to participants in the 2003 Punta Lobos Bioarchaeological Project: J. Marla Toyne, Sara Phillips, Sally Graver, Mellisa Lund Valle, and Lizbeth Escudero Lopez. Sara Phillips received an NSF Dissertation Improvement Grant for her doctoral research on Moche trauma patterns.

2 6 6   Bentley and Klaus

chapter ten

Reconsidering Retainers Identity, Death, and Sacrifice in High-Status Funerary Contexts on the North Coast of Peru Sylvia Bentley Behavioral Science Department, Utah Valley University

Haagen D. Klaus Department of Sociology and Anthropology, George Mason University

In many funerary traditions around the world, the tombs of high-status individuals are not only appointed with notable quantities or qualities of grave goods—they contain bodies of people as well (Baadsgaard et al. 2012; Campbell 2007; Parker Pearson 2000; Weiss-Krejci 2003). In the English-speaking Peruvianist tradition, such human remains are widely designated as “retainers.” In Spanish, acompanante (or companion) is commonly employed. Widely described as the wives, concubines, or servants of an elite principal personage, they have been enshrouded in functionalist terms, often cast as passive traveling companions or as a biological grave good. On the north coast of Peru, archaeologists have sometimes displayed something of a blind-faith presumption that tomb retainers were sacrificed in the process of preparing for the funeral of a highly ranked individual. Here, we examine the orthodoxy and evidence of retainer burials. Who, in fact, were these people? Were they human sacrifices? Exploring the answers to these questions helps take some first steps in developing a biocultural model of retainer burial via three lines of inquiry. First, a regional analysis examines available mortuary, skeletal, and taphonomic data associated with retainers in Moche, Sicán, Chimú, and Chimú-Inka high-status tombs (~AD 300–1552). Second, we integrate findings from

Reconsidering Retainers  2 67

biodistance analyses derived from dental phenotypes among a subset of Middle Sicán tombs to determine if kinship and identity played a role in who was selected to accompany these lords in death. Third, we examine the evidence and basis for inference surrounding the possibility that these individuals were ritualistically murdered as part of elite funerary rituals.

Ethnohistory and Archaeology of Retainer Burials The institution of Andean retainer sacrifice, sometimes referred to as necropampa immolation (Besom 2009), often defines a retainer as a person who attended to a living, high-status individual and provided various worldly services. Archaeological linkage of bodies in tombs to the ethnohistoric record owes much to second- and third-hand sources found in the work of 18 chroniclers writing between the years 1533 and 1644 (Besom 2009: table 3.3). In one prominent account, Betanzos (1996 [1557]) described those to be buried with their deceased lord as first being fêted and dressed in fine garments. After the victims were thoroughly intoxicated on chicha beer, an executioner strangled them from behind. Pizzaro (1921 [1571]) described the Inka “hanging” retainers upon Atahualpa’s death, while Cieza de León 1998 [1553] told of children who were buried alive with a dead curaca in Ecuador. The number of people reportedly buried with a dead lord varied, from one companion interred with a local lord, to thousands allegedly killed upon the death of the Inka emperor Huayna Capac (Cobo 1990 [1653]). Some sources comment on the identities of retainers and share a high degree of concordance in depicting them as wives, children, and court officials. Other sources designated retainers as the “old captains” and pages of an elite, while others yet were described as soldiers, attendants, and as the “beautiful women [who the lord] had loved” (Besom 2009:57). Most chroniclers understood that the sacrificed would follow a deceased lord into an afterworld to render services and provide company. Should a lord lack a basic necessity, such as food, drink, or attendants, the newly minted ancestor would seek supernatural vengeance upon the living (Bello Galloso 1897 [1582]). Further, Cieza de León (1998 [1553]) stated that burial with live victims bought great political prestige in death. Such conspicuous consumption no doubt conferred clout to living affines as well. Betanzos (1996 [1557]) and Zárate (1968 [1556]) noted that sacrifice played into principles of succession and maintenance of

2 6 8   Bentley and Klaus

royal orders. Competitors to the throne may have received the “honor” to be selected as a retainer. Betanzos also stated that royal Inka were motivated to include family and attendants in their burials to symbolically reproduce, legitimize, and perpetuate social structure and elite kinship ties to thwart potential political instability that surrounded death and rites of succession. These ethnohistoric accounts are provocative but must be consulted with caution on many levels. While potentially informative for the archaeological record, they are simultaneously incomplete, skewed, and depicted mostly by Europeans. Still, ethnohistory should not be summarily disregarded but instead carefully evaluated for potential heuristic value in analogies and interpretations. The bioarchaeology of retainers in Peru, and on the north coast in particular, remains nascent. In comparison to the focus placed on the principal personage of a tomb or description of other grave goods, there is less emphasis on these secondary personages beyond descriptions of age and sex. They are almost universally assumed to be sacrifices. They are not attributed agency, having been cast into the mold of superficial traveling companions akin to a human grave good. In one of the few detailed analyses of human remains in elite tombs on the north coast, Shimada and colleagues’ (2004) study of relationships among and between Middle Sicán principal and secondary personages shed light on the symbolic messages encoded by the living into symbolic “tomb choreography,” cultural constructions of human life histories, and sociopolitical organization. Similarly, Toyne’s (2002) detailed study of a group of “weaving women” at Túcume convincingly challenged previous assumptions, including their identity as members of an Inka aclla (groups of chosen women dedicated to the service in Inka Sun Temples). Such work demonstrates that if these archaeological data were pigeonholed into ethnohistoric expectations, we would scratch only the surface of more complex ancient realities. Currently, retainer burial practices involve more questions than answers. The narrative of this chapter involves three sections. First, we assemble a meta-analysis from the current body of literature involving published archaeological, taphonomic, and other details surrounding retainer burials on the north coast of Peru from Moche to Inka times. Second, we examine dental biodistance measurements of genetic relationships within a subset of Middle Sicán tombs to better assess the manners in which kinship and identity factored into these practices.

Reconsidering Retainers  26 9

Third, we evaluate the widespread affirmation that retainer burials were victims of ritual killing.

Exploring the Evidence of Retainer Burial, AD 300–1532 Retainers have been reported in nearly every major high-status funerary context on the north coast (table 10.1). Yet any meta-analysis involving these people will be uneven. A few excavations resulted in relatively detailed analyses, while a majority describe only the age and sex of individuals accompanying a principal personage. Retainers can be rather consistently identified due to several systematically patterned features, such as a noncentral or otherwise secondary spatial positioning within a tomb or the comparative lack of associated grave goods. Some displayed unusual or haphazard body and limb positions within otherwise highly regimented funerary settings. Taphonomic or archaeothanatological perspectives can be conservatively applied to descriptions or photos of skeletons in situ to consider depositional process and history. Age and sex variation may provide perspectives on the selectivity involving the choices of retainers. Much remains to be learned from the life histories of retainers before death vis-à-vis the bioarchaeological correlates of identity, social practices, ethnicity, and status (Gravlee 2009; Knudson and Stojanowski 2009; Krieger 2004).

Table 10.1. Summary of Retainer Burials on the North Coast of Peru Site Sipán

Tomb

N N SubN Retainers Adults adults

N Adult Males

N Adult Females

Tomb 1

8

7

1

3

4

Tomb 2

5

5

0

3

2

Tomb 3

1

1





1

Tomb 14

1

1





1

Tomb 7

1

1



1



Tomb 8

2

2







Tomb 11

1

1







Tomb 12

1

1







Huaca El Pueblo

Lord of Úcupe

4

4



2

2

San José de Moro

M-U41

5







5

27 0  Bentley and Klaus Table 10.1. Continued. Site

Dos Cabezas

El Brujo Complex

Tomb

N Retainers

N SubN Adults adults

N Adult N Adult Males Females

M-U1512

1



1





M-U1404

4

4







Mu- Priest

5

4

1

2

2

M-U1045

5



5





M-U615

58









Tomb A

0









Tomb B

3

3



2

1

Tomb 1

1



1



1

Tomb 2

2



2

1

1

Tomb 3

2



2





Huaca Cao Viejo Tomb 1A

10



2

4

4

Huaca Cao Viejo Tomb 2

13





1

1

Señora de Cao

4

4



4



Huacas de Moche Tomb 5

6









Huaca de la Cruz

4



1

1

2

Huaca Loro

Tomb 1 East Tomb

4



1

1

2

West Tomb

22

21

1



19











Illimo

Illimo Tomb 1

2

2





2

Chornancap

Tomb 3

2







1

Tomb 4

7

7





7

Chan Chan

Avispas Platform

93





8

85

Túcume

Huaca Larga TOTALS

19

17

2



17

296

85

20

33

160

– = Absent

The Moche Era The emergence of status differentiation (Elera 1998), precious-metal grave goods looted from the Chongoyape region (upper Lambayeque valley) (Lothrop 1941), and high-status tombs at Kuntur Wasi (Cajamarca) (Onuki and Inokuchi 2011) all indicate that traditions of elite burial in northern Peru may date back to circa 1000 BC. Yet retainers

Reconsidering Retainers  27 1

systematically appear first in Moche tombs (AD 100–850), with key contexts (highlighted below) proceeding geographically north to south. Sanctuario de Sipán, Huaca Rajada (Lambayeque Valley) To date, excavations have documented 16 elite Early-Middle Moche tombs at Sipán, described in detail elsewhere (Alva 1994, 2001, 2008; Alva and Donnan 1993; Verano 1997b). Tomb 1 (figure 10.1) contained the iconic Lord of Sipán, who evidently occupied the highest echelon of power. Buried with him in a large adobe funerary chamber were eight people. Of note was an adolescent female, wearing a copper headdress, whose degree of skeletal disarticulation was consistent with delayed burial (Verano 1995). To the east of the principal personage, a young male was found missing his left foot. Two adolescent females, also notably disarticulated, were stacked to the south of the principal person, with the woman on top in a face-down position. Atop the roof of the burial chamber, an adult male interpreted as the “guardian” of the tomb was buried with a copper disk (probably a shield) while both his feet were missing (Chero 2012). A meter above the chamber roof was a seated adult female, placed into a niche in the wall. The four individuals to the sides of the Lord of Sipán were interred in cane coffins, suggesting

10.1. A reconstruction of Sipán Tomb 1, containing the remains of the Lord of Sipán and seven other individuals. Courtesy of Museo Tumbas Reales de Sipán.

27 2   Bentley and Klaus

10.2. Examples of second-order tombs at Sipán: Tomb 10 (left) and Tomb 12 (right). Courtesy of Museo Tumbas Reales de Sipán.

differential status compared to those on the north and south, who were placed on their sides or facing down (Verano 1995, 1997b). Tomb 2 contained the remains of an older adult male, designated as The Priest, who was buried with five other people. To the east side of the official was a female, evidently in a sprawled, face-down posture (Verano 1995). Above the chamber, an adult male was placed in a cane coffin, wearing a copper semilunar crown and possessing a scepter, who was also missing his feet. Tomb 3 corresponded to the Old Lord of Sipán. Though its material opulence rivals Tomb 1, only a single person was buried with this lord. She was a young female between 16–18 years of age, lying facedown with her right arm extended above her head (Alva and Donnan 1993). Tomb 14 contained another Warrior-Priest roughly contemporaneous with Tomb 1 (Alva and Chero 2008) and also contained only one other person—a single young adult female. Burials of a lower nobility at Sipán included Tombs 5, 7, 8, 10, 11, 12, 15, and 16 (figure 10.2). Tombs 5, 9, 10, and 15 did not contain anyone in addition to the principal burial. Tomb 7 contained two people, a male and a female in cane coffins, and the male was missing his feet (Chero 2012). Tomb 8’s double burial contained high-status individuals

Reconsidering Retainers  27 3

accompanied by a pair of retainers. One of the principal burials and both companions again showed taphonomic signatures of delayed burial (Millaire 2002). In Tombs 11 and 12, a principal personage was accompanied by retainers, and in both cases their feet were missing. Huaca El Pueblo (Zaña Valley) At Huaca El Pueblo, Bourget and colleagues (2012) describe the Early Middle Moche Lord of Úcupe, a male in his mid-30s buried with multiple diadems, shields, and other sumptuous goods. The Lord of Úcupe was accompanied by an adult male to his right, an adolescent female directly beneath him, and another possible female individual directly beneath the principal personage of the tomb (Figure 10.3). The funeral bundle of the adolescent female contained the bones of a fetus, and she was evidently pregnant at the time of death (Klaus 2008b). A

b

c

10.3. The secondary personages in the tomb of the Lord of Úcupe. Images courtesy of Steve Bourget.

274   Bentley and Klaus

San José de Moro (Jequetepeque Valley) Surrounding the platform mound of La Capilla at the site of San José de Moro, multiple Late Moche and Transitional Period high-status tombs contained inferred retainers (Castillo 2011; Donnan and Castillo 1992, 1994; Millaire 2002). However, less is described about them in the literature regarding body positions, locations in the tombs, or bioarchaeological characteristics (but see Tomasto et al., chapter 11). Of particular note, the first elite tomb discovered, M-U41, contained a middle-aged female principal personage associated with paraphernalia indicating that she was a priestess. Five other females were placed around this woman (Donnan and Castillo 1994; Millaire 2002). Tomb M-U1404 was reopened and the principal personage removed. Subsequently, four other bodies were placed in the shaft of this boot-shaped tomb as it was being closed (Castillo 2011:81). Two inferred priestesses, two subadults, and a so-called guardian accompanied a Late Moche period male priest in Tomb M-U1727 (Castillo 2011). By the Early Transitional period, priestess burials continued at San José de Moro and included tombs such as M-U1045, which contained two female principal personages and a child accompanied by four other children and the legs of three adults. Most distinctive is M-U615: an open-access, collective tomb that contained two priestesses and the disarticulated remains of at least 58 additional people spread throughout four depositional levels—revealing a very complex depositional history and long-term funerary process (Rucabado 2008). Dos Cabezas (Jequetepeque Valley) On the Pacific shoreline of the Jequetepeque valley, Donnan (2007) documented five Late Early/Early Middle Moche elite tombs along the southwest corner of the principal adobe mound of Dos Cabezas. Tomb B featured the remains of two adult males and a female. The taphonomy of the remains is consistent with delayed primary burial (cf. Nelson 1998; Nelson and Castillo 1997), while a pair of lower human limbs and scatters of human bones were on the north and east sides of the tomb’s roof (Cordy-Collins 2009). Tomb 1 at Dos Cabezas contained two individuals. The principal adult male was accompanied by hundreds of items and an adolescent female. Donnan (2007) hypothesized she was a sacrifice due to her splayed position and the fact she was unclothed at the time of burial.

Reconsidering Retainers  27 5

Atop the roof of Tomb 2 was an adult camelid and the face down body of another woman, oriented atypically to the south and with her right forearm flexed under the body. Again, positioning and pairing of the woman with a camelid led Donnan (2007) to infer sacrificial death. Under the burial of the principal personage was an adolescent male accompanied by two stacks of gourd bowls and a copper nose clip. Tomb 3 contained three individuals. A child accompanied the adult male principal personage. The child displayed disarticulation of the right shoulder girdle and mandible consistent with delayed primary burial. Atop the roof of the chamber was an approximately 11-year-old child, but the skeleton was heavily disturbed due to the collapse of the chamber roof. Huaca Cao Vie jo (Chicama Valley) At the seaside El Brujo complex (Mujica 2007), large funerary chambers atop Huaca Cao Viejo contained Middle Moche tombs. Tomb 1A showed evidence of reopening in antiquity (Franco et al. 1998, 2001), with a fragmented principal personage and the disturbed remains of at least nine other individuals (three adult males, one adolescent male, three adult females, and two children) (Verano and Lombardi 1999). During the reopening event, another high-status woman was buried, accompanied by an adolescent female placed in a rare seated and flexed position. Tomb 2, also reopened, contained the remains of at least 13 distinct but disturbed individuals, some of which were likely retainers. Only two undisturbed individuals, an adult male and female, were found in the layer above the new roof (Millaire 2002). A high-status woman who was interred in a patio floor atop Huaca Cao Viejo has become known as the Señora de Cao (Franco 2008; Mujica 2007). The body of an adolescent was placed just outside her funerary bundle. The Señora was also surrounded by a quartet of ostensible companion tombs, which were self-contained adobe chambers. Each exhibited a less ornate context and contained a body in direct spatial and stratigraphic association with the principal burial. Huacas de Moche (Moche Valley) Since the work of Uhle (2014 [1913]) and Kroeber (1925), a large corpus of burials has emerged from Huacas de Moche (Donnan and Mackey 1978; Millaire 2002:82–96; and many chapters within the Investigaciones en Huaca de la Luna series edited by Uceda and Morales since 1997). Virtually all of these reported contexts, including high-status

27 6   Bentley and Klaus

interments, are single inhumations. Only a few exceptions exist in the urban zone’s Architectural Complex 35 (Tello and Delabarde 2008). There, six individuals surrounded the inferred principal personage of Tomb 5. Their facedown, splayed, haphazard, or incomplete remains led Tello and Delabarde (2008: fig. 183) to find these examples consistent with offerings. Interestingly, material correlates of social status indicate this context may be “middle class” (Chapdelaine 2001), leading to consideration of this unique setting as emulating high-status burial practices with potentially fluid boundaries between ritual and social class. Huaca de la Cruz (Virú Valley) As part of the Virú Valley Project, Strong and Evans (1952) uncovered a tomb at Huaca de la Cruz containing an adult male principal personage: the Warrior Priest of Virú. A child accompanied him inside the coffin. An adult female was at the foot of the coffin in a tightly flexed position, with “a blue and brown banded cotton sash of a much finer weave . . . around the neck, possibly indicating strangling” (Strong and Evans 1952:152). A second adult female was placed facedown in the opposite side of the grave. In the tomb fill, and closer to the surface, an adult male was present with his ankles and knees bound by cloth. The Middle Sicán Era Beginning in the early 1990s, Shimada and colleagues (Shimada 2014c) began sampling funerary contexts at Huaca Loro in the Pomac National Historic Sanctuary as part of their long-term study of the Middle Sicán (or Classic Lambayeque) culture (AD 900–1100) in the Lambayeque Valley Complex. At Huaca Loro, the East Tomb and West Tomb were elaborate and deep shaft constructions that contained elite personages and many additional individuals (Shimada et al., 2004). In the East Tomb, the principal personage was an older adult male covered in cinnabar and buried with nearly a ton of sumptuous grave goods. Above him was an individual 12–15 years of age in a semi-flexed position. A subadult around five years of age was found in a niche on the north wall of the chamber. Two adult women between 30–35 years old were buried with high-quality gold ornaments in the northwest corner of the tomb in what has been interpreted as a symbolic birthing scene (figure 10.4) (Shimada 1995).

Reconsidering Retainers  27 7

10.4. The two adult women on the floor of the Huaca Loro East Tomb buried with high-quality gold ornaments in the northwest corner of the tomb in what has been interpreted as a symbolic birthing scene. Drawing by Jahl Dulanto and Cesar Samillán and courtesy of Izumi Shimada.

In the West Tomb, the principal personage was a middle-aged male placed in a tomb-within-a-tomb extending 3 meters below an expansive antechamber. Here, 22 people (almost all young adult women) were found in the antechamber, distributed among 12 symmetrical pits on the north and south sides of the antechamber or in niches adjacent to the central chamber (figures 10.5 and 10.6). Each pit contained bodies in seated and flexed positions, and some were accorded formal grave goods such as ceramic vessels (Shimada et al. 2004). A long textile strip, wrapped around the waist of the principal personage, ascended up to the floor of the antechamber, where it came in contact with the women in the south side of the tomb. At the secondary center of Illimo 8km west of Sicán, Martínez (1997) encountered the pit tomb of a low-ranking elite from the Late Middle Sicán period. Two adult women in a seated and flexed position flanked this old adult male on his left and right sides. One of them was buried

27 8   Bentley and Klaus

10.5. Two of the young women (Burials 7 and 8) stuffed into an antechamber floor pit in the Huaca Loro West Tomb (see Fig. 5.6) during the Middle Sicán Period. Drawing by Izumi Shimada and Cesar Samillán and courtesy of Izumi Shimada.

wearing a labret. Her distinctive lip plug represented possible emulation of body adornment associated with the southern Ecuadorian Tallán culture by an apparently local woman (Klaus et al. 2004). After the fall of the Middle Sicán government and religious system, the Late Sicán (or Late Lambayeque) period represented an era of political reorganization from AD 1100–1375. Excavations at the monumental sector of Chornancap revealed three high-status tombs dating to the Late Sicán period (Wester 2012). Tomb 3 contained the body of a male holding tumi knives accompanied by a seated individual and an apparent secondary burial. Tomb 4 was more complex in its diverse range of ceramic offerings and precious-metal objects. This principal personage, the Priestess of Chornancap, was a middle-aged woman buried in full regalia. Five other seated individuals were placed symmetrically in two rows in front of her, along with another body lying on its side behind the Priestess. Between the five women, an isolated human skull was found along with the body

Reconsidering Retainers  27 9

of a camelid. Directly beneath Tomb 4 was Tomb 5, containing another elite personage accompanied by two additional people in a very poor state of preservation (Wester 2012; Klaus 2014d). Chimú and Inka Periods The Chimú empire (ca. AD 1100–1470) was ruled from the capital city of Chan Chan in the lower Moche valley. It has been a focus of intense looting since the Colonial period, resulting in the near total annihilation of an extraordinary and enormous funerary record. Most of the ciudadelas appear to have possessed an elite funerary platform that involved large, trench-like burial chambers containing Chimú lords. Burial chambers were flanked by as many as 40 cells presumably containing retainers. Evidence of an exceptionally large-scale pattern of burial tentatively emerged from Huaca Las Avispas in the Laberinto (or Fechech An) Compound. Pozorski (1971) and Conrad (1982) described what was initially identified as 300 females “stacked like cordwood” accompanying the burial of a Chimú king. Recent reanalysis of the remains by Nelson and Mackey (2011) using contemporary bioarchaeological methods revealed an MNI of 93 individuals, including some males. All shared a generally abnormal age-at-death distribution between 13–31 years of age. During the terminal late pre-Hispanic era, Túcume thrived as a regional Inka administration center. Excavations at the 700 meter–long Huaca Larga revealed three males in well preserved bundles in Huaca Larga Platform 2, Room 1, among them the hypothesized last Inka governor of Túcume (Narváez 1995a). In the adjacent Room 3, 19 individuals were buried in groups under the floors; this included five adult females between 31–40 years, nine adult females between 21–30 years, three females between 15–20 years, and two subadults of indeterminate sex (Toyne 2002). They were originally placed in seated and flexed positions. Designated as weaving women, they were accompanied by high-quality tools including thread, needles, and chalk (Narváez 1995b). Further, it was speculated they were members of an aclla, or a chosen elite group of weavers sacrificed in order to accompany these lords in death (Narváez 1995b). Later work by Toyne (2002) questioned this scenario: she tested and rejected some biocultural models of aclla identity and occupational specialization based on the evidence, which included patterns of health and musculoskeletal stress markers inconsistent with high status and habitual weaving activities.

2 8 0  Bentley and Klaus

Biodistance Analysis of Middle Sicán Tombs Theoretical advances in the bioarchaeology of identity and contextualized studies of biological distance and intra- and inter-cemetery kinship analyses (Knudson and Stojanowski 2009; Stojanowski and Schillaci 2007) have provided insights on mortuary patterns and social organization in antiquity, including on the north coast of Peru (Corruccini and Shimada 2002; Klaus 2008a; Sutter and Verano 2007). Patterns of biological relatedness can be useful to investigate biocultural relationships between principal and secondary personages. Dental biodistance data from Huaca Loro’s East Tomb and West Tomb, along with comparative non-elite samples, allow for tentative study of kinship patterns. A long history of scholarship has shown that tooth size and morphology are under strong genetic control (Hubbard et al. 2015; Hughes and Townsend 2013; Kieser 1990; Scott and Turner 1997). Dental phenotypes are often good if not strong reflections of underlying genotypic variation. Teeth allow for information and sample sizes to be maximized and reflect both maternal and paternal genetic signals. Theoretical and technical aspects of this analysis are described in detail by Klaus (2008a:403–407). In brief, Klaus used a Mitutoyo digital sliding caliper to measure maximum mesiodistal (tooth length) and buccolingual (tooth width) measurements from the maxillary and mandibular teeth from the individuals found in the Huaca Loro East and West Tombs, Huaca Sialupe, and Illimo. Nonmetric traits were scored according to Turner et al.’s (1991) Arizona State University Dental Anthropology Scoring System using plaster reference plaques (table 10.2). All statistical analyses were conducted in SPSS 21.0 (IBM Corp. 2012). These data were then subjected to multiple stages of pre-analysis data winnowing and preparation, removing any individuals or variables that possessed more than 20% missing data and variables that possessed zero variance. Any remaining missing data points were imputed using an expectation-maximization algorithm. To incorporate tooth size and nonmetric traits into a single dataset that incorporates the maximum amount of relevant phenotypic data, the data were normalized and converted into z-scores. Then, sexually dimorphic shape and size biases were controlled for by a c-transformation of the z-scores (see Howells 1989). Euclidian distance coefficients (d) are a non-parametric multivariate approach that is a straightforward technique to determine distances

Reconsidering Retainers  28 1 Table 10.2. Metric and Nonmetric Traits Used in Kinship Analysis among Adults from Middle Sicán High Status Tombs Maxillary Dentition

Mandibular Dentition

MD and BL measurements (mm) UI1*, MD and BL measurements (mm) LI1*, UC*, UP3*, UM1* LC*, LP3*, LM1* Labial Convexity, UI1*, UI2

Shoveling, LI1*, LI2, LC*

Shoveling, UI1*, UI2, UC*

Cusp number, LM1*–LM3

Double Shoveling, UI1, UI2, UC*

Anterior Fovea, LM1

Tuberculum Dentale, UI1*, UI2, UC*

Groove Pattern, LM1–LM3

Canine Mesial Ridge*

Distal Trigonid Crest, LM1–LM3

Canine Distal Accessory Ridge*

Deflecting Wrinkle, LM1

Metacone (Cusp 3),UM1*, UM2, UM3

Protostylid, LM1*, LM2, LM3

Hypocone (Cusp 4), UM1*, UM2, UM3 Hypoconulid (Cusp 5), LM1*, LM2, LM3 Metaconule (Cusp 5), UM1*, UM2, UM3

Entoconulid (Cusp 6), LM1*, LM2, LM3

Carabelli’s Trait, UM1*, UM2, UM3

Metaconulid (Cusp 7), LM1*, LM2, LM3

Parastyle, UM1*, UM2, UM3

Enamel extensions or pearling LP3, LP4, LM1, LM2, LM3

Enamel extensions or pearling UP3, UP4, UM1, UM2, UM3

Peg shaped incisor or molar

Peg shaped incisor or molar

Odontome, all premolars

Odontome, all premolars * Used in analysis

while also minimizing Bonferroni type I error (Corruccini and Shimada 2002). In Euclidian hyperdimensional n-space, the distance between points P (p1, p2, . . . pn) and Q (q1, q2, . . . qn) is determined by the square root of the measures’ average squared distance, or: d (p,q) = d (q,p) = √(q1–p1)2 +(q2–p2)2 + . . . + (qn–pn)2

Resultant linear measures of dental phenotypic distance should therefore reflect genetic distance; they are, of course, not necessarily directly proportional to genotype (Corruccini and Shimada 2002:117). Hierarchical cluster analysis was used to visually represent interindividual distances using the furthest neighbor or the complete linkage agglomerative method. It computes the distance between two clusters as

2 8 2   Bentley and Klaus

a

b 10.6. Burial 3 (A) and Burial 20 (B) in the Huaca Loro West Tomb. Drawings by Izumi Shimada and Cesar Samillán and courtesy of Izumi Shimada.

the maximum distance between a pair of clusters (see Everett et al. 2001). Complete linkage tends to highlight naturally distinct clusters in datasets, and it is generally immune to chaining, a form of erroneous clustering. To check the “goodness-of-fit” and mathematical validity of these clusters, they were validated using a two-step cluster silhouette measure of cohesion and separation (Rousseeuw 1987; Rousseeuw et al. 2006). This is a standard validation method in R and SPSS (Bacher et al. 2005). Each cluster is represented by a so-called silhouette that

Reconsidering Retainers  28 3

measures the degree of confidence of the clustering algorithm in terms of Euclidian distance. Silhouette values range on an interval between –1.0 to 1.0. Randomly structured or spurious clusters range from –1.0 to 0.0. Clusters considered of a “fair” quality fall between 0.0 and 0.5. Confident, non-random, and appropriately structured clusters are considered “good” and range between 0.5 and 1.0. The resultant Middle Sicán clusters (figure 10.6) possessed a silhouette value of 0.50, at the breakpoint between “fair” and “good” cluster

10.7. Hierarchical cluster analysis visually representing estimated genetic distances between individuals in Middle Sicán tombs and individuals from the nearby sites of Illimo and Huaca Sialupe. PP = principal personage; B = burial; ET = East Tomb; WT = West Tomb.

2 8 4   Bentley and Klaus

quality. Thus, the clusters appear to represent non-random patterns, and it is possible to say that this graphical representation of Euclidian distances demonstrates the nature of the relationships within the dataset. These patterns of phenetic similarity hold interesting implications. Comparisons were made between principal individuals and retainers and a small (N = 22) but diverse sample of the local, non-elite population derived from cemeteries at Illimo and Huaca Sialupe (Klaus et al. 2004). Two large clusters emerge. One contains the lords of the East Tomb and West Tomb and their inferred kin. In particular, relatively close distances were calculated between the principal personage and Burial 3 (the upright women in the birthing scene) in the East Tomb. Individuals from non-elite sites in the Lambayeque region dominated the other cluster. In the West Tomb, individuals on the south side of the antechamber appear relatively closely related to the lord as well as Burial 5 in Niche 5 (northwest corner). However, most of the young women on the north side of the tomb appear not to have shared as many genes in common with the lord and display greater affinity to members of the local non-elite population.

Physical Evidence of Violent Retainer Death As demonstrated by various authors in this book, there are multiple potential physical reflections of sacrifice, which Verano defines as the “intentional offering of human life” (2001a:167). Intentional violent death may leave unambiguous skeletal evidence such as sharp force trauma, blunt force trauma, stab wounds, the presence of ligatures, and even body positioning reflective of live burial. This section of the chapter is somewhat short. Osteological study of the better-preserved human remains in virtually every north coast tomb does not report a single unambiguous cut mark or other form of perimortem trauma consistent with intentional violent death. Only a few exceptions exist. Tomasto et al. (chapter 11) describe a basicranial ring fracture in an inferred retainer at San José de Moro. Preserved ligatures were found around retainers’ necks in three settings (Huaca de la Cruz, El Brujo, and Manchan, a Chimú administrative center in the Casma valley) (Strong and Evans 1952; Verano 2001a). In other contexts (i.e., footless bodies at Sipán or Huaca Loro 2006, Trench 2, Burial 6), the evidence exists in a kind of grey zone. Context may strongly suggest

Reconsidering Retainers  28 5

sacrifice or mutilation, but preservation is suboptimal, which prevents decisive evaluation of violent trauma.

Discussion The burial of additional people with a principal individual was an enduring element of high-status funerary ritual among the complex societies of north coast Peru. Our regional analysis identified 274 retainers accompanying 31 principal personages in Moche, Sicán, Chimú, and Inka settings. Complexities of Identity The identities of retainers and expressions of selectivity emerge first on demographic levels. Of these 274 cases, demographic information is inconsistently published. From the available data, we identified at least 20 subadults, 85 adults of indeterminate or unstated age or sex, 33 adult males, and 160 adult females. Children, representing 7.3% of the sample, are most infrequently represented, whereas adult females are most common (58.4% of the sample). The high percentage of females is strongly skewed due to the large number of women represented in the single context of Las Avispas. Still, it may be that, following the Moche, increasing relative proportional representation of women in contexts such as the Huaca Loro West Tomb, Chornancap, Las Avispas, and Huaca Larga represent a cultural focus in retainer burial selection. Biodistance analyses confer biocultural perspectives to questions of retainer identity. In the Middle Sicán Huaca Loro East and West Tombs, some retainers appear to be closely related kin, while others are more distantly associated. Comparison with other non-elite Middle Sicán individuals from around the valley confers a broader perspective. Here, some of the associated women seem drawn from rather distinct and homogenous groups of leaders. This is consistent with previous studies of mtDNA variation, other dental phenotypes, and FST analysis reported by Corruccini and Shimada (2002), Klaus et al. (in press), and Shimada et al. (2004). Others were related to the local commoner population composed mostly by inferred ethnically Muchik peoples. Identity likely played a role in the design and choreography of the West Tomb, with the close female affines of the lord buried to his south, plus women ostensibly drawn from the local population to his north. Analysis of health

2 8 6   Bentley and Klaus

patterns by Farnum (2002) and Klaus et al. (in press) indicates the south side women shared with the principal personage very good skeletal and oral health, suggestive of similar social experiences and access to sufficient nutritional resources since childhood. This contrasts with the health of the women on the north side of the West Tomb. Their higher prevalence of pathological conditions suggests inferior nutrition and more chronic stress throughout their lives, which links them to inferred non-elite groups. Shimada et al. (2004) explored the potential significance of tomb organization vis-à-vis Middle Sicán multiethnicity and political symbolism, but Ramírez (2004) raised another issue. North coast ethnohistory indicates that important indices of the status and power of a lord were reflected in the number of his followers and his ability to offer hospitality to anyone who appeared before his court. Colonial documents often referred to a lord’s female attendants as “widows,” who may have also included orphans, captive females, or persons entering into the service of a lord (i.e., household labor). The relative kinship diversity seen in these Moche and Middle Sicán contexts (some secondary personages as inferred close kin, while others were far less closely related) is reasonably consistent with early Colonial accounts of the composition of a royal court. Still, caution must surround these estimated kinship patterns. Withinlineage marriages would explain the biological homogeneity documented among Middle Sicán elites and raise the possibility of endogamy to the point of sibling marriage, perhaps to consolidate and protect power within elite lineages (Klaus et al., in press). Thus, a closely related female could also have been a wife. These are not mutually exclusive categories. The nature and identity of children and males in tombs is even harder to pin down, though maternally related males in the case of Sipán (Shimada et al. 2005b) indicate their presence in a tomb could have correlated to an avunculate-based system of descent. The bodies of children were placed in close or even intimate physical proximity to a lord, as seen in settings such as Sipán Tomb 1, Dos Cabezas Tomb 3, Huaca de la Cruz, and the Huaca Loro East and West Tombs. This is a very intriguing pattern that we do not fully understand. At least in the Huaca Loro East and West Tombs, those subadults share close biological ties to the principal personages. We might speculate they were favored young male relatives whose lives mattered more in terms of service to the lord in death. Or, within an avunculate system of descent, is it

Reconsidering Retainers  2 87

possible these were the children incapable of lineal political succession? Even more challenging is inferring anything concrete about unrelated male retainers. Their identity as court functionaries is within the realm of probability if there were any similarities to Inka practices, but definitive demonstration of their status or roles may never be fully resolved. Retainers as Sacrifice Victims? Ethnohistoric perspectives undeniably point to intentional killing producing retainer burials. Physical evidence of perimortem trauma is absent in nearly every setting, failing to meet the empirical and evidentiary standards for ruling on violent death in a modern court of law. Yet other observations lead to a series of three complications. First, ethnohistory and direct archaeological evidence, such as that at Huaca de la Cruz (Strong and Evans 1952), El Brujo, Manchan (Verano 2001a), Huambacho (Chicoine, chapter 4), and Pachacamac (Uhle 1991 [1903]), reveals that ligature was an established method of ritual killing. Strangulation does not typically produce skeletal trauma. The hyoid bone is spared from fracture in almost all cases of death by ligature or manual strangulation (Ubelaker 1992). It is also impossible to discount the chance that secondary personages were poisoned, drugged, drowned, or killed using some other technique that did not leave evidence in skeletal tissue. Telltale signs of live burial, such as blindfolds or limb positioning produced via binding are frequently absent, but in at least one case at Huaca de La Cruz lower limb binding was identified. All of this indicates that funerary sacrifices were set apart from blood sacrifices. Due to this distinctive kind of killing, the absence of evidence does not necessarily equate to evidence of absence in most kinds of north coast contexts. Retainers, by their nature, may always involve ambiguities surrounding their lives and deaths. Sacrifice may be conservatively (if not reasonably) inferred. Such interpretations must be drawn from judiciously documented archaeological contexts via contextual, probabilistic, and evidence-based inference—while also embracing the inherent ambiguity and the limits of available evidence (i.e., Eeckhout and Owens’s [2008] concept of the potentially sacrificed individual). For example, the sex and age distributions in tombs at Sipán, Sicán, and Las Avispas are similarly skewed toward young adult females. Such demographic skewedness is a strong mismatch with a naturally occurring death assemblage (Coale

2 8 8   Bentley and Klaus

and Demeny 1966). These patterns seem to reflect multiple cultural selective filters and even expressions of value judgments. The fact that younger females are chosen as retainers in so many diverse settings probably reflects the importance of emic age categories, conceptions of personhood, perceptions of the differing values placed on particular people’s lives, and female gender roles in life and death. Such demographic discrepancies call strong attention to the presence of cultural filters that operated nowhere else in these societies. One could still fall victim to confirmation bias to conclude that such selectivity represented unnatural deaths (see below), so the strength of such an argument must hinge on the concurrent evaluation of other lines of archaeological context. Second, inference of sacrificial death requires careful observation of in situ skeletal articulation. Evidence of prolonged primary burial (Huchet and Greenberg 2010; Millaire 2002; Nelson 1998; Shimada et al. 2015; Verano 1997b) is known from many funerary contexts on the north and central coast, spanning from subtle to extensive disarticulation. Prolonged burial appears to be the case with Sipán Tomb 1 Retainers 3, 4, and 7 (adolescent females) (Verano 1997b). The missing feet of the various poorly preserved guardians at Sipán were not necessarily removed by violent, mutilatory amputation as pondered by Alva (2001). Instead, feet may have been manually broken off from a brittle, desiccated corpse before interment. In the tomb of the Lord of Úcupe, the shroud-wrapped female was associated with the jumbled bones of a preterm fetus by her knees. The best explanation for this observation was that the woman’s body began to decompose before burial. Abdominal bloating and rupture during putrefaction probably expulsed the fetus. Exposed fetal connective tissues, with their lower mass, decomposed at a faster rate than the mother’s body (Duday 2009). Later movement of the woman’s bundle into its final resting place further jumbled and shifted her bones along with the fetal remains. Taphonomy of several women in the Huaca Loro West Tomb demonstrated entomological activity, various states of disarticulation, and missing terminal finger and toe phalanges (Shimada et al. 2004). Together, these data indicate such bodies were delayed interments. The point we are driving at is that, within many popular and scholarly visions of elite burial, timing is implicitly invoked to make a case for retainer sacrifice via mass synchronous deaths. In other words, a lord dies and wives and officials are sacrificed for the funeral. Evidence of prolonged primary burial reveals near-certain non-contemporaneity between some principal

Reconsidering Retainers  28 9

personages and members of their entourage. Contemporaneous death should never be assumed. Some retainers may have died months, years, or even decades before the funeral took place. Some retainers may have indeed experienced natural deaths. If this is the case, does the abnormally narrow age and sex range suggest another cultural filter, one where specific people’s bodies were amassed as they randomly died, as they may have been already destined to become a retainer in death? In any case, taphonomy points to scenarios involving accumulation of some corpses and planning for a lord’s funeral for a substantial amount of time before his death in a long-term process of retainer death/sacrifice and curation. They are not necessarily reactionary ritual killings fulfilling pressing needs of a funeral. Retainer curation also raises another issue to come full circle, back to the critique that opened this chapter. Might have a humanized identity slipped away from curated retainers who were indeed transformed in objectified offerings, grave goods, or sacred relics in these such instances? Third, ethnohistory offers tantalizing observations that could suggest some retainers were not sacrifices, but rather were produced in acts of suicide or auto-sacrifice. Betanzos (1985 [1557]) described numerous people, so impatient to join the dead Inka emperor Tupac Yupanki, that they strangled themselves—men used rope while women committed suicide using their belts. Cobo (1990 [1653]) related that upon the death of Huayna Capac many people gave up their lives, of their own accord, to join their emperor. Retainers committing ritual suicide—and not having violent external agency imposed upon them—is a striking prospect. Identification of ritual suicide in these settings is, of course, probably archaeologically impossible. Still, it remains a distinct possibility in these ancient realities.

Conclusion Retainers in elite tombs blur the lines between archaeological typologies of sacrifice, and assessment of their identities and deaths involves formidable interpretive challenges. Biodistance analysis helps clarify the complex social and kinship relationships in these contexts while simultaneously posing new questions. Contextual clues suggest many retainers suffered unnatural deaths but were not mutilated as victims of blood sacrifice. Yet others may not have been victims at all but committed ritual suicide. Some evidently died long before the funeral.

2 9 0  Bentley and Klaus

The additional bodies accompanying a principal personage in highstatus tombs may always reside in the space between empirical certainty and evidence-based inference. Still, they represent a relatively unexplored area of the mortuary and bioarchaeological records. Even still within the published literature, a next step could involve formal multivariate analyses of burial treatment, demography, and other attributes along with new interpretations of the nature of retainers as passive offerings versus active agents within the symbolic choreography of elite funerals. Biogeochemical assessment of the geographical origins of secondary personages will provide another important development. Next-generation DNA sequencing of ancient genomes also holds great promise to definitively address many of the questions posed here. While retainer burials may never be free of ambiguity, increasing focus and study of these individuals will help close various gaps of knowledge in the archaeology of Andean mortuary patterns. Acknowledgments We thank Walter Alva, Bruno Alva, Steve Bourget, Luis Chero, Christopher Donnan, Kimberly Jones, Andrew Nelson, Izumi Shimada, and Carlos Wester for their generous access to samples, sharing of data, images, and insights. J. Marla Toyne and four anonymous reviewers provided excellent editorial insights and corrections.

Human Sacrifice  2 9 1

chapter eleven

Human Sacrifice A View from San José de Moro Elsa Tomasto-cagigao Pontificia Universidad Católica del Perú; Programa Arqueológico San José de Moro

Mellisa Lund Programa Arqueológico San José de Moro, Pontifica Universidad Católica del Peru

Luis Jaime Castillo Pontificia Universidad Católica del Perú, Programa Arqueológico San José de Moro

Lars Fehren-Schmitz University of California, Santa Cruz

Archaeological thinking about sacrifice on the north coast of Peru has been deeply influenced by the evidence found at Huaca de la Luna and Dos Cabezas (Cordy-Collins 2001a; Verano 2001a, 2001b, 2008a). However, as chapters in this volume demonstrate, other forms of sacrifice were practiced in the region involving ways of killing that may have not produced skeletal trauma. In this chapter, we focus on two funerary contexts from the site of San José de Moro in the Jequetepeque valley—one dating to the Late Moche period (AD 700–850), and the other to the subsequent Late Transitional Period (AD 1200–1300). Both contexts contained several complete and incomplete skeletons and many commingled bones. As has been previously demonstrated, the Moche practiced complex funerary rituals that included extensive manipulation of human remains (Nelson 1998; Nelson and Castillo 1997; Verano 1997a), and the two funerary contexts described in this chapter are no exception. Such practices may complicate the identification of human sacrifice that may be associated with these tombs. Nevertheless, a rich

2 9 2   Tomasto et al.

body of contextual information, as well as carefully defined evidence of perimortem modification of bones in these tombs, allow us to propose that unique forms of human sacrifice may have been practiced at this key site on the north coast of Peru.

Frameworks and Settings Two basic questions arise when we address human sacrifice within archaeology. First, how do we define it? Second, how can we identify its practice in the archaeological record? In other words, how do we differentiate sacrifice from, for example, secondary offerings of human remains or the collection and curation of human body parts (Verano 2001a)? Regarding the first question, the definition of human sacrifice is twofold: the first element involves ritual relationships with the realm of the sacred, as the etymology of that word implies. But this is a partial definition of the nature of sacrifice, because, for example, offerings of human bones removed from a reopened burial are also sacred, yet they are not a product of sacrifice. This realization leads to the second part of the definition: the necessity of the offering (or some constituent element of it) to be destroyed or killed (Hubert and Mauss 1964 [1899]) in an attempt to please the sacred or fulfill some kind of sacred obligation. This twofold definition implies that the identification of human sacrifice in archaeological contexts must be based first on the demonstration of a sacred context and, second, that an individual’s death was unnatural. Only rarely are human remains associated with both conditions in the archaeological record. When preservation is good, the latter criterion can be easily fulfilled. Nevertheless, demonstrating the cause of death in skeletonized remains is very difficult and usually can be done only when the events that led to death involved violence that damaged bone. Owing to a degree of preservation often absent in the archaeological record, the discovery of sacrificial victims at El Brujo allowed for the identification of ropes tied around the necks of several females (Verano 2001a:170), which gives support to the hypothesis of sacrifice in Moche elite burials. Strangulation as a method of ritual killing has been little explored on the north coast of Peru, even though iconographic data suggest that it was practiced (see Chicoine, chapter 4). The sacrifices of Huaca de la Luna and Dos Cabezas (Cordy-Collins 2001a; Verano 2001a, 2008a) are among the few cases in which human sacrifices can be demonstrated in skeletonized remains.

Human Sacrifice  2 9 3

In this way, can sacrifice be definitively identified in cases of poor preservation or in the absence of skeletal trauma resulting from unnatural death? Verano (2001a) thinks that the context where the remains are found can provide important clues for sacrifice. Could bodies with unusual positions, locations, or biological profiles be considered indirect evidence of some kind of unnatural death? Some individuals in the high-status burials at Sipán or the tomb of the lord of Huaca de la Cruz fit this model. Particularly in the second case, such criteria work rather well to infer sacrifice in that the simultaneous death of several children of similar ages is improbable (Bentley and Klaus, chapter 10; Mogrovejo 1995). In this chapter, we go further. In combining taphonomic and bioarchaeological analyses, we argue there is evidence that supports human sacrifice from skeletonized remains even though there is little skeletal trauma. San José de Moro San José de Moro is the largest cemetery under investigation on the north coast of Peru, in terms of both chronological span and overall size. San José de Moro is located in the northern Jequetepeque valley, close to the modern town of Chepén and to other Moche sites such as Cerro Chepén and San Ildefonso (Castillo 2012). The long occupational history of San José de Moro showcases the intricate social and cultural processes that shaped the north coast during the second half of Moche cultural history and beyond (Castillo 2012; Castillo et al. 2008). Twenty-two years of investigation have revealed that the site was not solely a cemetery; it also contained evidence of complex ceremonial activities that produced an intense occupation. Tombs cannot be understood in isolation of the ritual contexts within which they came into being (Delibes and Barragán 2008). The site was occupied during five consecutive periods (figure 11.1). During the Middle Moche period (AD 500–700), irrigated lands were expanded to the northern part of the valley, and San José de Moro was occupied for the first time. In this period, San José de Moro was a peripheral site in relation to centers such as Pacatnamú and was clearly in the sphere of the northern Moche Middle period phenomenon that had its most dominant locus at Sipán. With few exceptions (Ruiz 2008), large or complex Middle Moche burials are absent at San José de Moro. Ceramics during this period are exclusively Moche and surprisingly poor in quality when compared to the Moche ceramic traditions of the south.

11.1. Cultural chronology of the occupation at San José de Moro. Image by the Proyecto Arqueológico San José de Moro.

2 9 4   Tomasto et al.

Human Sacrifice  2 9 5

The Late Moche period (AD 700–850) signaled a departure from Middle Moche traditions, as new ceramics styles appeared in association with larger tombs. Boot-shaped shaft tombs contained individuals of intermediate social status. Larger chamber tombs became favored by elites and contained greater quantities of grave goods and multiple individuals as retainers for a main occupant. A third type of tomb involved shallow and narrow burial pits that seem to have contained commoner burials with limited types of offerings (Castillo and Uceda 2008:722–723). Ceramic styles reveal a complex panorama of cultural interactions. The most emblematic was the Late Moche fineline painting style used to decorate elaborate stirrup-spout bottles (Castillo 2012), while several foreign ceramic designs appear in the larger chamber tombs. The inclusion of these artifacts signals the end of Moche isolation in this region. In this setting, the Moche did not seem to be passive in the face of the Wari Empire. In fact, it may have been the Moche who lured the Wari into the Jequetepeque region (Castillo 2000b). As an immediate consequence, a new hybrid tradition—the Late Moche polychrome style that synthesizes Moche and Wari styles—was created at San José de Moro. The Transitional Period (AD 850–1000) emerged after the collapse of the Moche polities. During this period, coastal traditions expressed an unprecedented degree of highland influences, and the ceramics of the Wari viñaque style, as well as multiple Cajamarca substyles, became prominent in San José de Moro burials. The typical boot-shaped tombs disappeared, and chamber as well as pit tombs predominated, with the latter containing individuals of intermediate social status. The Lambayeque (or Sicán) period (AD 1000–1200) saw the cessation of all eclecticism and innovation, placing the Jequetepeque valley back into a north coast cultural track. Extensive evidence of typical Lambayeque artifacts and funerary contexts are found only in pit tombs. There is little evidence of a dense occupation, and San José de Moro was no longer the center of religious life in the valley. The subsequent Chimú occupation (AD 1200–1400), signaled the beginning of the abandonment of the site, and it became a work-related area in the periphery of the larger regional center at the adjoining Algarrobal de Moro. Finally, even though ethnohistorical sources mention a strong Inka control of the Jequetepeque valley, only one Inka style ceramic sherd has been found at the site in association with a Chimú chicha (maize beer) production facility.

2 9 6   Tomasto et al.

In sum, for over 1,000 years (AD 500–1500), San José de Moro was one of the most important ceremonial centers and cemeteries in the northern Moche region (Castillo and Donnan 1994b), and during Late Moche times it was the dominant ceremonial and pilgrimage center in the valley. The Moche and Transitional peoples returned to the site, possibly on a calendrical schedule, to celebrate their ancestors, to feast and pray, to participate in ritual reenactments, and to interact with peoples of other Moche towns and hamlets. Funerary Practices at San José de Moro Although we could define three distinctive tomb types (pits, boots, and chambers), a universal funerary pattern could not be clearly discerned at the site. Some shared funerary rules or behaviors are perceived in the orientation of tombs, the position of bodies and artifacts, and to a lesser degree the spatial organization of tombs in alignments and clusters (Castillo 2012). But each complex tomb at San José de Moro seems to be more of a unique biographical statement of who was buried, rather than a strict observation of idealized norms. Nevertheless, and for the sake of simplicity, it is instructive to examine typical tomb types found at San José de Moro. Pit tombs were the most common and appeared in all occupational periods. They were usually a simple, narrow, and elongated hole in the ground that contained the deceased and a few offerings. In contrast to the other two types of tombs, a great diversity of cardinal orientations characterized pit tombs (Donley 2008). In Moche times, pit tombs were reserved for low-status individuals, although some examples exist of exceptional ceramic bottles used as grave goods. Most pit tombs contained the bodies of commoner women and young children closely associated with chicha kitchens and brewing areas. The humble pit tombs gained a certain degree of prominence during the Transitional Period, when larger pits sometimes contained individuals of intermediate status (Castillo and Rengifo 2008). During the Lambayeque and Chimú occupations, pit tombs were the only type of tomb at San José de Moro, but by then the site was no longer the cemetery of choice for high-status people. Boot-shaped shaft tombs appear at San José de Moro with its earliest occupants and disappear with the collapse of the Moche polities. This kind of tomb is almost exclusively found in Jequetepeque. It may have originated from the interaction between the local Moche and the

Human Sacrifice  2 97

Moche-Vicús peoples in the northern Piura valley (Donnan 2007, 2011). During the Middle Moche period, boot-shaped tombs were dominant, and they contained lower-status individuals buried with simple funerary offerings (Del Carpio 2008). The only exception to this pattern was a cluster of larger and lavish Middle Moche tombs found in the northern sector of San José de Moro (Ruiz 2008). During the Late Moche period, boot-shaped tombs were quite common and contained individuals of intermediate status. Grave goods were more numerous in these tombs and frequently included Late Moche fineline and polychrome ceramics. It is with boot-shaped tombs that clear differences began to manifest between the main occupants, for whom the tomb was built and fitted, and other individuals who were included in the tombs, probably as retainers or sacrificial victims. The main occupants were located at the center of the tombs and often inside cane coffins, extended on their backs, and oriented north-south. Additional individuals were found stacked atop each other, either in prone, flexed, or otherwise aberrant positions. Frequently, these individuals have been called “guardians,” inferring that they were placed in the tombs so as to safeguard the main occupant. In at least two cases, the main occupant was a child, while the “guardian” was an adult male. Also, some boot-shaped shaft tombs have included multiple individuals that could correspond to discrete family groups (Castillo 2011, 2012). Chamber tombs were the largest funerary contexts found at San José de Moro but were also the most rare (Castillo 2012). In contrast to funerary chambers found at other Moche sites such as Sipán and Huaca Cao Viejo, where tombs were built inside adobe brick mounds, chamber tombs at San José de Moro were built directly in the ground by digging a rectangular hole, often several meters deep. In some cases, the excavation had a ramp at one end connecting the ground surface with the floor of the chamber. Inside the rectangular pit, a four-walled room was loosely built from adobe bricks, often without using mortar. Most chambers included rectangular niches in three of the walls. The chambers varied tremendously in size, from as small as 1.5 by 2.5 meters to as large as 7 by 7 meters. Once the chamber was completed, it was roofed using the same techniques that the Moche used to build houses, creating an equivalent “house for the dead.” In all cases, there is no reason to think that the chambers existed prior to the burials but, rather, that they were constructed as part of these funerary ceremonies. Whether they were built in advance of the arrival of the occupant and offerings is unknown.

2 9 8   Tomasto et al.

Once ready, the chambers were filled with artifacts and bodies. The coffin of the main occupant was positioned most prominently, usually at the center of the tomb or on top of a slightly raised step. The coffin itself seems to have been a special artifact, decorated with a large funerary mask on one end, sheet-metal arms and legs, or metal bands or decorated tassels on the sides. Decorated coffins could have been paraded around San José de Moro as part of funerary ceremonies and would have been the most visible aspect of the funerary persona. Retainer burials, “guardians,” and possible sacrificial victims were placed next to the principal interment, followed by other kinds of offerings. These surrounded the coffin and could include thousands of ceramic, metal, wood, textile, and stone items. Usually the niches built into the tomb walls contained some of the finest and most unusual objects. After all the elements had been put in the tomb, it was partially filled with clean sand, the roof was finished, and additional offerings were placed on top of the roof. Finally, the shaft was filled. Transitional Period chambers and some of the Late Moche chamber tombs showed evidence that they were visited and reopened after they were closed and that some degree of alteration had occurred. In some cases, additional artifacts and even individuals were added; in other cases, grave goods and even the body of the main occupant were apparently removed. One of the most interesting aspects of these tombs is that they include several individuals (frequently incomplete or disarticulated) in addition to the main occupant. This was particularly true in the case of five chamber tombs that contained adult females identified as Moche priestesses associated with the Sacrifice Ceremony (Castillo 2005, Donnan and Castillo 1992). Their tombs included several individuals, and in some cases significant quantities of disarticulated bones (particularly skulls and long bones) had been placed around the coffin or in the niches. Other individuals usually were placed in atypical positions, with extended or flexed limbs, facing down or contorted, such as in the examples of bootshaped tomb guardians. Several individuals were represented by entire or almost complete skeletons.

Suspected Sacrifices in Late Moche and Transitional Tombs With more than 400 documented burials pertaining to the Middle and Late Moche periods and 200 additional burials dating from the subsequent Transitional period, one could expect to find evidence of human

Human Sacrifice  2 9 9

sacrifice associated with the elite burials. The funerary contexts where human sacrifices most probably occurred at San José de Moro were those in which priestesses were buried (Castillo 2005). Here, we examine the evidence from M-U1525 (Late Moche period) and M-U1221 (Late Transitional Period). Both contain the remains of high-status female principal personages closely related to the priestess cult. Combining taphonomic data recorded in the field with careful examination of skeletal remains in the laboratory, we identified an extended and very complex sequence of funerary rituals practiced at San José de Moro that included a high degree of manipulation of the bodies. These practices, previously described by Nelson (1998), Nelson and Castillo (1997), and Verano (1997a), complicate identification of human sacrifices, because they could easily be confused or conflated with secondary offerings of human remains. Nevertheless, the rich contextual information, as well as the existence of some evidence of perimortem modification of bones, allows us to propose that some form of previously unidentified human sacrifice was likely practiced at this important site on the north coast. Burial M-U1525 This burial, dating to the Late Moche period, was a large chamber tomb that contained the remains of at least 14 individuals associated with opulent offerings (table 11.1). As in most funerary structures of this type, it was built with adobe bricks at the bottom of a deep shaft. Niches were present in three walls, and the roof was built with algarrobo logs, adobe bricks, and reed mats that were supported by posts. The entrance to the chamber, located to the west, was an independent structure formed by three walls set in the shape of a U with a step leading to the chamber (figure 11.2a). The chamber itself was divided into two sections by a bench, creating a raised space to the east and an antechamber to the west. Various levels of sediments recorded in the fill reveal that a liquid was poured in repeatedly, perhaps during reopening rituals. Also, a small wall enclosing the entrance was built, apparently to protect it against blowing sand while it remained opened. Grave goods were distributed throughout the fill, in the niches, and inside the chamber itself. More than 3,000 miniature vessels of unbaked clay, known as crisoles, were located in groups, throughout the chamber and in the fill, on fragments of big pitchers, or inside pots and bowls. There were also several skeletons of camelids, both complete

3 00  Tomasto et al. Table 11.1. Burial M-U1525 Skeletal Summary Individual Sex

Age

Location

Comments

1

F

25–35

Above offerings

Primary burial

2

F

Old adult

Bench, inside coffin

Secondary burial

3

Indeterminate

15

Entrance of chamber

Secondary burial

4

F

Old adult

On bench

Secondary burial

5

M

20–30

Chamber

Secondary burial

6

M?

Middle adult

Chamber

Secondary burial

7

F

Middle adult

Chamber

Secondary burial

8

F

40–50

Corner of chamber

Primary burial, atypical orientation

9

Indeterminate

Child

Chamber

Disarticulated, very incomplete

10

Indeterminate

Child

Chamber

Disarticulated, very incomplete

11

Indeterminate

Child

Chamber

Disarticulated, very incomplete

12

Indeterminate

Indeterminate

Niche

Only skull

13

Indeterminate

Indeterminate

Niche

Only skull

14

Indeterminate

Indeterminate

Niche

Only skull

and in pieces, eight architectural models made from unbaked clay, and dozens of pitchers, plates, and bottles, with some showing depictions of mythological Moche beings. Truly remarkable elements included metal objects of priestess paraphernalia: a goblet, two plumes, and part of a triangular headdress. Two coffins were badly preserved, but both had a funerary mask and copper sheets representing a priestess, although only one maintained its original form. At least 14 other individuals with differing degrees of skeletal completeness and articulation were buried in this tomb (table 11.1). All were inside the chamber, and most were extended with the head toward the south. The last individual interred, probably during the final sealing of the tomb, was Individual 1, a woman 25–35 years old (figure 11.2a). She wore several necklaces and bracelets and was laid above the level where all other individuals and most grave goods were located. Some of her

Human Sacrifice  3 01

vertebrae and ribs and both feet were disarticulated. Nevertheless, the range of movement of these bones was inside the expected norm when a body decomposes in a primary context (Duday 1997). Toward the east, on the bench, two more skeletons were found. Individual 2, an elderly woman (figure 11.2b), was inside a coffin. Her vertebral column and pelvis were articulated, but all other bones were displaced. The skull and most bones of the right upper limb were absent. Evidently, this woman was completely decomposed, and the coffin had undergone extensive movement or manipulation preceding placement in this tomb. Perhaps she was taken from another tomb or was removed and replaced in the same tomb. As part of the activities of relocating or reopening, some bones were also removed from this coffin. On the left side of the coffin were the very incomplete remains of Individual 4, another elderly woman (figure 11.2b). She was in a prone position, and her right upper limb and both lower legs were absent. Her left upper limb and cervical vertebrae were also disarticulated, suggesting that she A

b

c

d

11.2. Burial M-U1525: (A) upper level with entrance; (B) floor of the tomb; (C) the carefully arranged pelvis of Individual 3; and (D) bones of children among the bones of Individual 6. Photos by the Proyecto Arqueológico San José de Moro.

3 02   Tomasto et al.

received the same treatment as the woman inside the coffin, although in this case the container for the body was not preserved. At the side of the bench were the commingled remains of three adults and three children. Only Individual 7, a middle-aged woman located farther away from the bench, was easily individualized (figure 11.2b). Her skeleton was almost complete and in anatomical position, except for the feet, vertebrae, and ribs. The degree of movement of these bones does not match with the typical movements expected in primary contexts (Duday 1997), suggesting that this is also a secondary burial. However, the final deposition occurred when the body was in the initial phases of decomposition, unlike Individuals 2 and 4. The other adult skeletons, Individuals 5 and 6 (figure 11.2b), were located closer to the bench and corresponded to a 20- to 30-year-old male and to a middle-aged adult, probably another male. Both skeletons were disarticulated, commingled, and incomplete, reflecting careful and intentional manipulation of some bones. For example, the os coxae were placed one above the other, and the ribs were placed in groups. A copper knife was found among the vertebrae and ribs of one skeleton. Below the bones of the adults, the dispersed and incomplete remains of at least three children were found (Individuals 9, 10, and 11; figure 11.2d [arrows]). Two more skeletons were found toward the entrance of the burial chamber. Individual 8, a 40- to 50-year-old woman (figure 11.2b), was at the right side of the entrance near the corner. She was perfectly articulated, but unlike the other individuals her head was oriented toward the west and one of her legs was flexed. The other skeleton, Individual 3, was a nearly complete 15-year-old adolescent placed across the entrance (figure 11.2b). In this case, the absence of some vertebrae and the careful relocation of the os coxae (figure 11.2c) is evidence for a secondary burial, with a degree of disarticulation similar to that of Individual 7. In addition to the 11 individuals just described, three crania and other human bones from adults and children were found in two niches and inside the chamber. It was not possible to determine their origins. From the previous description, we can conclude that the individuals buried in this chamber died at different moments in time. The individuals showing the highest degree of disarticulation, and hence probably the greatest degree of decomposition at the time of burial, were the two old women placed on the bench and the two men and three children placed in close proximity. The woman resting next to them and the adolescent in the entrance showed less extensive disarticulation, but they were also

Human Sacrifice  3 03

secondary burials. Finally, only the woman lying near the corner of the chamber (Individual 8) and the woman found on top of the assemblage (Individual 1) were well articulated, indicating they died shortly before the final closing of the tomb. None of the bones presented cut marks or other signs of perimortem violence. Nevertheless, the preservation of the bones was poor. Traces of violent death may have once been present but not preserved. Burial M-U1525 could be hastily interpreted as containing one principal individual accompanied by several others who were sacrificed to honor the first. The clear predominance of female skeletons, including two elderly women, suggests that this is not a familial group. But at the same time, this inference is tempered by the lack of skeletal trauma and the evidence of reopenings and secondary treatments, revealing that not all the individuals died at the same time. The reopening of the tomb and movement of bodies are evidence of extended rituals that may be part of the veneration for a principal individual involving several potential sacrificial events. The unusual placement of the two women who died a short time before the final closure of the chamber—with atypical orientation and position in one case, and on top of the offerings in the other—may indicate that they were sacrificed. We will return to this point in the discussion. Burial M-U1221 Burial M-U1221 dates from the Late Transitional Period. This pit tomb contained comparatively few objects inside when compared to M-U1525. Yet this context similarly included several individuals with differing degrees of anatomical integrity and articulation (table 11.2), revealing a highly complex sequence of ritual events that may have also involved human sacrifice. The tomb was prepared by digging a shaft oriented north-south. At some point the excavation ceased, and two spaces were created: an entrance to the north, and a deeper funerary chamber to the south. Here, the individuals and offerings were placed. Two postholes found near the east wall of the shaft suggest that at some moment the tomb had a roof, perhaps because the tomb remained open for some time to carry out the complex rituals described below. Grave goods included plates, jars, an anthropomorphic vessel, miniature vessels, piruros, a mortar and pestle, minerals, two flutes made from

3 04   Tomasto et al. Table 11.2. Burial M-U1221 Skeletal Summary

A

Individual Sex

Age

Location

1

Indeterminate

Child

Entrance to chamber Intentionally manipulated

2

Indeterminate

Child

Left, above Ind. 4

Intentionally manipulated

3

M

35–45

Center

Bones in situ

4

F

Old adult

Left, above Ind. 6

Bones in situ, whistle in ischiopubic region

5

M

25–30

Right, above Ind. 7

Intentionally manipulated

6

F

30–35

Bottom, left

Intentionally manipulated

7

F

Old adult

Bottom, right

Bones in situ

b

Comments

c

11.3. Burial M-U1221: (A) Individuals 6 and 7, the initial occupants of the tomb (the skull of Individual 7 was picked up earlier in the excavation); (B) Individuals 4 and 5 arranged over the initial occupants; and (C) bed of bones in the middle of the tomb. Photos by the Proyecto Arqueológico San José de Moro.

Human Sacrifice  3 05

11.4. Burial M-U1221 with all its occupants exposed. Photos by the Proyecto Arqueológico San José de Moro.

bird bone, and polished pebbles with anthropomorphic and zoomorphic shapes. Several of these objects are similar to those used today in shamanic mesas, and it is possible that the individuals interred in this burial were involved in such activities (Castillo and Rengifo 2008a). The first occupants of the tomb were Individuals 6 and 7: two women placed in extended positions, one beside the other, with their heads toward the south end of the tomb. The woman to the left of the tomb (figure 11.3a [Individual 6]) was 30–35 years old. The woman on the right (figure 11.3a [Individual 7]) was an elderly woman. Over her right shoulder were several objects, including beads, shells, polished pebbles, and a copper knife. Both skeletons were articulated but incomplete. The absence of small bones such as those of the feet or some vertebrae can be explained by fluctuations in the water table that are frequent in this region. Flooding of the space could produce flotation of small bones and their movement within the tomb. Indeed, several small bones were found dispersed (see Backo, chapter 3, for a discussion of water transport of bones). The displacement of the skull and first cervical vertebrae of the older woman (figure 11.4 [Individual 7]) can also be explained via this mechanism. Nevertheless, the body of Individual 6 (the younger woman) was evidently manipulated. Her head was missing, and although three skulls with a biological profile matching this woman’s skeleton were found in different locations inside the tomb (figure 11.4), none of these could be

3 06   Tomasto et al.

matched to Individual 6. Also, eight superior left ribs, the left shoulder girdle, and the left humerus were missing. However, the forearm and hand preserved their anatomical positions, as if they were still articulated to the humerus, and the hand was grasping a flute made of bird bone. The missing bones were found within an assemblage of bones (figure 11.3c, center) that served as a sort of bed on top of which Individual 3 (figure 11.4) was lying. The os coxae of Individual 6 were also moved and rested in an inverted position on the lower thoracic vertebrae. As with the left forearm, the sacrum and legs maintained an intact anatomical position. The sternum was found next to the right os coxae. No cut marks or other signs of violence were found in the bones. Clearly, the bones of Individual 6 were moved when the body was in an advanced state of decomposition or perhaps skeletonized, probably during a tomb reopening. A finding that reinforces this hypothesis is an elongated greenish-black stain that surrounds the distal left radius and ulna. X-ray fluorescence analysis1 showed high levels of silver and copper in the stain. This suggests that the corpse was likely adorned with a metal bracelet that underwent some degree of oxidation and stained the bone but was later removed by the living. Apparently, this tomb was re-entered several times, and during these events two more bodies were added. Above Individual 6 (the young woman whose body had been manipulated) was placed Individual 4, an elderly woman (figure 11.3b, figure 11.3c). Those who arranged the bodies took special care in matching the pelvic bones of both women. A ceramic whistle was found on the sacrum and crossing the ischiopubic region of the older woman. This object was probably introduced when soft tissue was still present. Atop this perfectly articulated skeleton, a second, partially incomplete but otherwise articulated skeleton was found, corresponding to a child 1–2 years old (figure 11.4 [Individual 2]). Individuals 4 or 7 were not likely the mother of this child because of their advanced age. The younger woman (Individual 6) is also rejected as the possible mother since her mitochondrial haplogroup did not match the child (table 11.3). As with other individuals in this tomb, the left arm of Individual 2 was missing, as well as the left leg and right lower leg. Some of the bones were found dispersed in the center of the tomb within the assemblage of bones underneath Individual 3. It is important to mention that this child showed reactive bone on endocranial surfaces, suggesting the presence of a chronic illness at the time of death.

Human Sacrifice  3 07 Table 11.3. Mitochondrial DNA Variation, Burial M-U1221 Individual

Haplogroup

HVR I motive (sequence data-haplotype)

2

B2

16189C, 16217C

3

D1

16093C, 16223T, 16325C, 16362C

6

D1

16093C, 16223T, 16325C, 16362C

On the opposite side of the tomb and above Individual 7 rested the incomplete skeleton of Individual 5, a 25- to 30-year-old man (figure 11.3b). Only the skull, cervical and thoracic vertebrae, all right ribs, the three lower left ribs, manubrium, the right shoulder girdle, and the right arm were found in situ. Again, cut marks were not observed on any bones, and most of the missing bones were found in the central assemblage alongside the displaced bones of Individuals 6 and 2. Therefore, we think that Individual 5 was also complete when he was buried and that his bones were removed during a subsequent re-entering of the tomb when his remains were skeletonized. The final event in this sequence of rituals appears to have been the disposal of Individual 3 (figure 11.4), a 35- to 45-year-old man adorned with large ceramic ear spools. His body was laid out at the center of the tomb, between Individuals 6 and 4 to the left and 7 and 5 to the right, and on top of an assemblage of disarticulated bones (figure 11.3c, center). This group of disarticulated remains pertained to Individuals 2, 5, and 6 and also included camelid foot elements and other animal bones that did show cut marks. Offerings, including six human crania, were arranged above and toward the feet of the individuals. Finally, toward the north of the shaft and at the entrance to the funerary chamber rested Individual 1, another incomplete child, 1–2 years old (figure 11.4). The right torso, arm, and femur were missing and were not found inside the tomb. As in previously described cases, the rest of the skeleton was in anatomical position and did not show cut marks. Considering the similarity with the other skeletons, it is likely that this individual was also manipulated after decomposition. However, it is not possible to know at what moment this body was deposited during the prolonged use of this tomb. As previously mentioned, none of the incomplete skeletons or disarticulated human bones showed cut marks or injuries similar to those found in Moche captive sacrifice contexts (Verano and Phillips, chapter 9).

3 08   Tomasto et al.

11.5. Cranium OH07: detail of the plastic deformation at the union, superior part of the occipital bone. Also note the division of the skull into two large cranial fragments and the plastic deformation of the superior occipital bone. Photo by Lucía Patsias.

11.6. Detail of Cranium OH07’s basicranial ring fracture at the right occipital condyle and the fracture at the petrous portion of the temporal bone. Photo by Lucía Patsias.

However, three crania presented evidence of traumatic lesions and thermal exposure, indicating perimortem injuries and postmortem manipulation. One cranium, designated OH07 (figure 11.4), included only the vault and cranial base and corresponded to a middle-aged woman. A transverse fracture divided the cranium into two halves. The regular

Human Sacrifice  3 09

borders, areas of plastic deformation, and associated radial fractures suggest it was a perimortem fracture (figure 11.5). Another perimortem fracture was identified at the foramen magnum by the superior border of the right occipital condyle (figure 11.6). This fracture is characteristic of so-called ring fractures to the base of the skull. Basicranial ring fractures can be produced in two ways: when the foramen magnum punches down into the vertebral column by a direct blow to the top of the head, or when the vertebral column punches upward into the base of the skull when an individual falls from a height and lands on the feet or in a sitting position (Galloway 1999). Another perimortem fracture in the same cranium, involving complete transverse breakage of the petrous portion of the right temporal bone, is likely related to the ring fracture. Petrous portion fractures are uncommon because this bone is massive and well buttressed against external force. According to Galloway (1999), direct blows to the head producing longitudinal fractures of the parietal or temporal bones may propagate as linear fractures into the petrous portion of the temporal bone. Even though a fall from a height can produce this damage, in this case all contextual information indicates that this individual suffered one or more very strong blows that caused the separation of the cranial vault from its base, broke the petrous portion, and produced a basicranial ring fracture, as well as possible other fractures on the face that were not preserved. The implement used to produce such traumatic damage must have been massive. A second cranium, designated OH03 (figure 11.4), corresponded to an adolescent and was complete. It possessed an almost circular hole, approximately 5 millimeters in diameter, in the middle of the frontal bone (figure 11.7). The inner surface of this perforation possessed a clearly beveled edge, whereas the external border showed a small, depressed area of plastic deformation on the edge (figure 11.8). No radial fractures were observed. These characteristics are similar to those usually found in trophy heads, particularly those of the Nasca culture. Verano (2003b) argues that a frontal bone perforation (done to pass a cord and hang the skull) and damage to the base of the skull (to accomplish brain removal) are two characteristics of trophy heads. In the present case, there is no alteration to the base of the skull, and the borders of the perforation do not show the use wear produced by movement of cord against bone in the Nasca specimens. Nevertheless, it is clear that the perforation was made when the bone was fresh—and

31 0  Tomasto et al.

11.7. Ectocranial view of the perforation in the squamous portion of the frontal bone of Cranium OH03. Photo by Lucía Patsias.

11.8. Detail of the internal beveling, Cranium OH03. Photo by Lucía Patsias.

it seems intentional. This skull may have undergone initial preparation as a trophy but was never finished. The third cranium, OH05 (figure 11.4), corresponded to another female individual. It was darkened at the inferior occipital and temporal bones around the base of the cranium (figure 11.9). This coloration is more accentuated on the inner surface. It is consistent with effects of postmortem thermal alteration and exposure to fire.

Human Sacrifice  31 1

11.9. Burned occipital bone fragments in OH05. Photo by Lucía Patsias.

Discussion: Human Sacrifice at San José De Moro? Burials M-U1221 and M-U1525 have much in common. Apparently both remained open for a time or were reopened several times, as they were the final repository of various individuals that died at different moments. Many of the bodies were intentionally manipulated, and bones were moved or removed. Also, as discussed above, even though both burials contained the remains of men, women, and children, they were not likely family groups. On the contrary, females demographically dominated both tombs, and among them two elderly individuals stand out in each burial because they received distinctive treatments. This is not surprising at San José de Moro, where many powerful women linked to the priestess cult were buried. Perhaps in life these elderly women served as the priestess that frequently is depicted surrounded by an entourage of women (Castillo and Rengifo 2008b). In Burial M-U1525, both old women were found in the main part of the chamber above the bench. In Burial M-U1221, even though they were not in the central part of the tomb, both old women were complete and in anatomical position, whereas their companions were manipulated and some of their bones ended up in something akin to a bunk bed for another individual and were commingled with animal bones.

31 2   Tomasto et al.

In Burial M-U1525, it is clear that at least the old woman inside the coffin was the principal personage. Although there was no evidence of violence in the other skeletons, the women designated as Individuals 1 and 8 (table 11.1) showed atypical dispositions: one was among the offerings, and the other was lying haphazardly in one corner. Both individuals died a short time before the final closure of the chamber. The combination of this evidence suggests that these two atypical individuals may have been sacrificed to honor the main individual. While trauma is absent, this does not eliminate the possibility of a violent end, as various ways of killing that do not affect bone have been reported on the north coast. For example, several females found at the El Brujo Complex and in a high-status Sicán male burial were strangled with ropes found still in situ (Verano 2001a:170; also Chicoine, chapter 4). Unfortunately, the often poor preservation of organic material such as cordage at San José de Moro does not allow us to detect this type of evidence. In Burial M-U1221, the principal personage seems to be the adult male laid atop the others. But in this case, it is hard to state that his companions were sacrificed to honor him, because it is clear that all died before him. Still, there are hints of possible sacrifice. First, in the two groups of skeletons located on both sides of the main individual, there were recurrent oppositions: old/young, complete/incomplete. The manipulation of the bones of the incomplete individuals, and their subsequent placement alongside animal bones showing cut marks to shape the base for another interment, suggest that this could be an act of symbolic violation in the sense proposed by Duncan (2005). If so, a difference may exist between the four initial occupants of the tomb, with the two old women as principal personages in two events involving the sacrifice of two younger individuals that were further disarticulated and used as sacred objects. There are also two children who apparently received a treatment similar to the young adults, suggesting that these children may also have been sacrificed. At this point an analogy can be established with the sacrifice rituals of Moche prisoners in which, after sacrifice, some bodies were dismembered in order to divide the “human ritual vessel” into sacred objects suitable for use as dedicatory offerings (Hill 2003). In Burial M-U1221, dismemberment was not produced by cutting fresh bodies but rather by manipulating the decomposing remains of sacrifices. The three crania showing traces of heating, probable massive blunt force trauma, and evidence of an unfinished or rushed production of a

Human Sacrifice  31 3

trophy head may represent other possible sacrifices in this tomb. Particularly, the fractures present in OH07 are strongly suggestive of a very violent death followed by deposition as an offering. The additional modification of these remains suggests a high degree of variability in how remains were treated. Finally, among the objects associated in this tomb were a whistle and two flutes. Such musical instruments have been directly related to the performance of Moche sacrifice, specifically in that “the act of whistling is linked with the crucial moment of a ritual, just before the offering of a child, the sacrifice of a victim or the burial itself” (Bourget 2001a:113). Returning to the approach described at the beginning of this chapter, we proposed two conditions to identify human sacrifices in archaeological contexts: sacredness, and unnatural death in the context of a ritual event. Funerary contexts are usually the result of sacred action, and in the present case there are no reasons to argue that the funerary contexts described here fail to match the first criterion. The second criterion is more challenging to demonstrate, as there is no physical evidence of violent death with the exception of cranium OH07. At San José de Moro, this standard of proof is even harder to attain not only due to the often poor skeletal preservation but also because complex funerary rituals involving manipulation of bones may have removed key evidence and result in confusion with secondary offerings. Nevertheless, the rich contextual information that we presented, showing atypical treatments, and the analogies with other Moche contexts of sacrifice strongly suggest that this practice was also performed at San José de Moro during the Late Moche and Transitional Periods.

Conclusion In this chapter, we have shown that a careful analysis of human remains that brings together many lines of evidence can aid in the tentative identification of sacrifices in the absence of direct evidence of violent death. Many previous investigations have demonstrated that human sacrifices were common on the north coast, were frequently undertaken to honor members of the elite, and included methods of killing that did not damage the bones (Verano 2001a; also see Bentley and Klaus, chapter 10; Chicoine, chapter 4). In the Moche and Transitional contexts described here, a detailed analysis of the skeletons that identified recurrent biological profiles, atypical funerary treatments, and postmortem manipulation

314   Tomasto et al.

allowed the detection of possible differences among the interred individuals that, particularly in Tomb M-U1221, might have remained unseen due to the complex extended rituals. Among the treatments of the possible retainers, their manipulation, the disposition of the human remains as offerings, and two examples of perimortem lesions suggest violent death and some kind of violation (Duncan 2005). Therefore, they may represent sacrifices. The death of two individuals a short time before the final closure of a tomb that either remained open for an extended period or was reopened many times is suggestive of sacrifice, just as is their atypical body positioning. In conclusion, the poor preservation and ritual complexity have converged at San José Moro to prevent the conclusive identification of human sacrifices, although there are many clues suggesting that they probably did occur and may have taken on some unique forms involving thermal alterations, postmortem processing, and blunt force cranial trauma. The continuing investigations at this and other archaeological sites on the north coast of Peru will provide further evidence to test the hypothesis proposed here. Acknowledgments The excavations at San José Moro were carried out with the financial support of the Pontificia Universidad Católica del Perú, National Geographic Society, The Backus Foundation, Patronato de Las Huacas del Valle de Moche, and other institutions and individuals to all of whom we are sincerely grateful. Burial M-U1525 was excavated by Cecilia Mauricio, and Burial M-U1221 was excavated by Carlos Rengifo. The skeletons were analyzed in Lima, kindly supported by Consuelo Cuadrado. The preparation of bones for analysis was carried out by students of the PASJM Bioarchaeology Field School and Lucía Durand. The English manuscript was kindly revised by Ann Peters. We appreciate the invaluable comments made by the three anonymous peer reviewers, the volume editors, and especially John Verano, as always.

Life Histories of Sacrificed Camelids  319

chapter twelve

Life Histories of Sacrificed Camelids from Huancaco (Virú Valley) Paul Szpak Department of Anthropology, The University of Western Ontario, London, ON, Canada

Jean-François Millaire Department of Anthropology, The University of Western Ontario, London, ON, Canada

Christine White Department of Anthropology, The University of Western Ontario, London, ON, Canada

Steve Bourget Department of Anthropology, University of Montreal, QC, Canada

Fred Longstaffe Department of Earth Sciences, The University of Western Ontario, London, ON, Canada

The practice of sacrifice or ritual killing in ancient central Andean societies has become a topic of increasingly focused study in archaeology and physical anthropology. Much of this research has examined the killing of human victims in a range of geographic and temporal contexts (Andru-shko et al. 2011; Bourget 2001b, 2006; Gaither et al. 2008; Klaus et al. 2010; Sutter and Cortez 2005; Toyne 2008; Tung and Knudson 2010; Verano 2008a; Wilson et al. 2007). There has also been a recent surge in bioarchaeological research, which has allowed investigators to reconstruct the circumstances surrounding the death of these individuals in great detail (Klaus et al. 2010; Toyne 2008; Verano 2008b). Isotopic studies have been employed to some extent for sacrificial victims with the aim of better understanding the life histories of these individuals

3 2 0  Szpak et al.

(Fernández et al. 1999; Wilson et al. 2007) or to specifically address their geographic origin (Andrushko et al. 2011; Toyne et al. 2014; Tung and Knudson 2010; Turner et al. 2013). The purpose of this research is to utilize an isotopic approach to reconstruct the life histories of sacrificed llamas from Huancaco in the Virú valley of northern Peru. This line of investigation says less about how these individual animals came to die and more about how they lived—although these two factors may very well be interlinked. Such information is valuable for understanding sacrifice or ritual killing within its larger economic, social, and ritual contexts.

South American Camelids The South American camelids consist of two wild (vicuña and guanaco) and two domestic (llama [Lama glama] and alpaca [Vicugna pacos]) species. The llama and alpaca were the only large mammalian species to be domesticated in the Western Hemisphere and were prominent in the economic, social, and ritual spheres of numerous Andean societies (Bonavia 2008; Dillehay and Núñez 1988; Dransart 2002; Miller and Burger 1995; Murra 1965; Shimada and Shimada 1985). From an economic perspective, llamas were utilized for meat and as beasts of burden, being instrumental in long-distance exchange networks (Browman 1990; Dillehay and Núñez 1988; Nielsen 2001). Both llamas and alpacas were utilized as fiber-producers, although the finer fleece of the alpaca made them prized for this purpose (Murra 1962; Orlove 1977). Much of what is known about camelid herding practices in the Andes comes from ethnographic and ethnohistoric accounts of herders in high altitude areas (>3,800 meters above sea level) (Flores-Ochoa 1979; McCorkle 1987; Murra 1965, 1980; Orlove 1977; Tomka 1992). Some of the more detailed descriptions of historic camelid management practices come from Murra (1965, 1980), who describes Inka herding during the Late Horizon Period (AD 1476–1532), primarily in southern Peru. Within this context, there were distinct herds owned by the state, the church (so-called shrine herds), and the community. The sizes of these herds varied greatly from one region to another, and in some cases certain types of herds (especially shrine herds) appear to have been absent in some areas. In many cases, community herds were of much smaller size than the church or state herds, and several accounts suggest that community-owned animals were not contributed

Life Histories of Sacrificed Camelids  32 1

as sacrifice or tribute (Polo de Ondegardo 1571). There are records that state-owned and shrine herds had access to specifically demarcated pastures that could not be grazed by outside herds (Murra 1965). Moreover, there are numerous indications that the physical attributes of the animals (color in particular) were of some significance with respect to selection for particular types of sacrifice (Murra 1965). It is problematic to extrapolate ethnohistoric accounts of Inka rule too broadly, but these data do suggest the possibility that sacrificial animals may have been raised differently from non-sacrificial animals and perhaps even perceived in a wholly different manner than animals destined to become food or supply wool. These differences may have been dietary and consistently applied throughout the lives of the animals, or they may have occurred close to the times of the animals’ deaths, as in the case of force-feeding llamas chicha prior to sacrifice (Webster 1972). More generally than this, we may expect that sacrificial animals were physically segregated from other animals and fed a more specific or less diverse diet in accordance with what was believed to be ideal food—or at least ideal food for an animal destined for sacrifice. Analogous cases for such practices come from Vanuatu in Melanesia, where certain types of pigs with ceremonial importance in prestige-related feasting are fed special diets and kept in houses (Blackwood 1981; Funabiki 1981; Jolly 1984). Such a pattern may be expected to manifest itself in a low amount of isotopic variability for a group of sacrificial animals compared to non-sacrificial animals. An important question in Andean prehistory concerns whether or not camelids were raised locally in the coastal river valleys or were raised exclusively in high-altitude environments and imported to the coast (Shimada and Shimada 1985). Following European colonization, camelid pastoralism has been predominantly limited to the high-altitude pastures of the Andes, and this is reflected in ethnohistoric and ethnographic accounts (e.g., Flores-Ochoa 1979; Murra 1965; Tomka 1992). Recent isotopic studies of coastal camelids have, however, provided evidence that camelids were indeed raised on the coast beginning at least during the Early Intermediate Period (Szpak et al. 2014) and through to the Middle Horizon (Dufour et al. 2014; Szpak et al. 2014). On the basis of substantially higher amounts of inter-individual isotopic variation and inconsistent amounts of inter-individual variation, Szpak et al. (2014) have argued that coastal camelid husbandry consisted primarily of numerous small groups of animals kept within or around

3 2 2   Szpak et al.

urban environments by individual households. This pattern stands in contrast to the much larger herds described for high-altitude pasturelands. Despite having local camelid herds on the coast (likely llamas rather than alpacas due to their tolerance of a wider variety of environmental conditions [Topic et al. 1987]), fine wool was still being imported from the highlands to the Virú valley during the Early Intermediate Period (Szpak et al. 2015). A coherent picture of the origins and nature of coastal camelid husbandry is still emerging, but at this point evidence suggests experimentation with keeping animals locally during the Early Horizon (Szpak et al., in press), before the practice was more deeply established during the Early Intermediate period (Szpak et al. 2014).

Animal Sacrifice Animal sacrifice has been discussed in a diverse array of ethnographic and ethnohistoric texts (Abbink 2003; Christman 2008; Cole 1997; Evans-Pritchard 1956; Hasu 2009; Insoll 2010; Middleton 1987; Pluskowski 2012; Rahimi 2004; Sprenger 2005; Wiget and Balalaeva 2001). Typically, these discussions have focused on the ritual aspects of animal sacrifice and, in turn, reflect the symbolic importance of certain animals or animal species and the larger symbolic significance of these events. There is an array of views concerning exactly what constitutes the sacrifice of an animal rather than the killing of an animal under “ordinary” circumstances (reviewed by Russell 2012: 88–138). On the basis of ethnographic evidence of hunter-gatherers and pastoralists, the death of many if not most animals may be highly ritualized. For instance, in some hunting cultures, the hunt consists of a series of events that incorporate significant spiritual aspects, and the killing of the animal by the hunter is an event with implicit or explicit ritual aspects within a larger set of beliefs about humans, animals, and the environment (Bird-David 1990; Brown 1984; Carneiro 1970; Nadasdy 2007; Willerslev 2004, 2007). Similarly, among some East African pastoralists, every animal slaughter—including those destined solely to produce food—is a highly ritualized event. By most conventional definitions (and as defined by these peoples themselves), it is a sacrifice (Ryan et al. 2000). Additionally, many instances occur among herders in which sacrificial materials are drawn from live animals without the death of the animal, as in the case of ritual bleeding (Abbink 2003; Dransart 2002; Evans-Pritchard 1940).

Life Histories of Sacrificed Camelids  323

Therefore, in an attempt to simplify “animal sacrifice,” we focus on animals that are killed under highly ritualized contexts in which the ritualized component is not incidental but is foregrounded. Within archaeological contexts, these types of animal sacrifices may be identified through contextual associations (Cross 2011; Goepfert 2012; Hamilakis and Konsolaki 2004; Lucas and McGovern 2007; Verhoeven 2002), such as in conspicuous displays of animal remains (Lucas and McGovern 2007), the deposition of complete animal carcasses in association with human burials (Cross 2011; Goepfert 2012; Losey et al. 2013), in ritual contexts (Green 1992), or the burning of animal remains (Hamilakis and Konsolaki 2004; Isaakidou et al. 2002). In some cases, where there is no clear evidence of butchery (or even slaughter), it is typically inferred on a variety of evidence that animals deposited in such contexts represent sacrifices (Croft 2003; Goepfert 2012; Russell and During 2006; Sandweiss and Wing 1997; Wheeler et al. 1995; Yuan and Flad 2005). Regardless of the manner in which such animals died, the deposition of entire animal carcasses that may have produced significant quantities of materials in life (milk, wool, traction, fertilizer, transport) and in death (meat, hides, wool, other raw materials) clearly represents a sacrifice in the classical sense (Hubert and Mauss 1964 [1898]). Drawing primarily on ethnographic literature, two fairly common aspects of animal sacrifice are: (1) the use of domestic animals (largely or exclusively); and (2) the incorporation of the animal sacrifice within a gift economy with ancestors or supernatural entities. The use of domestic animals is significant in the sense that sacrifice (rather than simply offering) implies the giving of something that is owned by the donor (Firth 1963a, 1963b; Ingold 1987, 1994; Testart 2006). This relationship better characterizes herders and their animals than it does animals that are hunted (Ingold 1994). The notion of sacrificed animals as part of a supernatural gift economy has deep roots in anthropology (Tylor 1889) and, to some extent, is overly simplistic in that it fails to account for the larger symbolic or metaphorical aspects of the ritual event itself (Lienhardt 1961; Valeri 1994) and of the performative nature of the sacrifice (Cole 1997; Hoskins 1993). A more reasonable approach recognizes a complex set of relations that are historically situated in the sacrificial event among several parties: sacrificers and sponsors, ritual practitioners and specialists, victims (in this case animals), spirits or ancestors, and the human audience (Hasu 2009; Hoskins 1993). There may be overlap among these positions in any given ritual.

3 2 4   Szpak et al.

The distinction between wild and domestic animals in sacrificial events is of further significance. This is by virtue of the different nature of human-animal relations that characterize hunters and those involved in animal husbandry respectively. Specifically, in the case of wild animals, individuals selected for sacrifice are unlikely to be familiar with some members of the community (Knight 2005, 2012). Conversely, for domestic animals, it is more likely that some individuals are, to at least some extent, known to the community. Perhaps this line may be blurred in cases where wild animals are captured and held for appreciable periods of time prior to sacrifice (Batchelor 1908; de Sales 1980; Trigger 1990), but these distinct cases likely occupy some liminal state between wild and domestic and are, of course, exceptions to the broader pattern. The significant point here is that the lives of the animals that were sacrificed were shared with members of a human community, with a high degree of mutual familiarity between human and animal. This is particularly noteworthy since, in many instances in sacrifice, the animal is thought to be, to at least some extent, a substitute or proxy for the human sponsor providing the animal for sacrifice (Hamayon 1990; Hasu 2009; Rasmussen 2002). The implication here is that particular animals are directly associated with particular people or groups of people. This directly ties in to the potential importance of prestige and the competitive nature of sacrificial ritual, where individuals, families, or kin groups provide victims for sacrifice, rather than communities (Gibson 1988; Heesterman 1993). More generally, the sacrifice is situated within a larger spiritual, social, economic, and political context, within which the ritual practitioners conducting the sacrifice represent only one segment. To more completely understand the nature of this event, it is also necessary to examine the lives or biographies of the sacrificed animals, as well as the sacrificial event itself. Such examination is facilitated by an array of techniques utilized in modern bioarchaeology and zooarchaeology.

Camelid Sacrifice in the Peruvian Andes The sacrifice or ritualized killing of animals played an important role in ceremonial activity in the ancient Andes. On the north coast, Moche iconography depicts the hunting of deer and sea lions in highly ritualized manners (Benson 1995; Bourget 2006; Donnan 1997a). While these activities involved the killing of wild animals and are thus distinct from those

Life Histories of Sacrificed Camelids  32 5

involving domestic animals for reasons discussed above, if we accept the premise that hunters are engaged in some type of gift economy with the animals that they take and the supernatural (Nadasdy 2007), then these hunts could also be viewed as sacrifices. Similarly, articulated skeletons of numerous wild species (particularly birds) have been recovered and interpreted as offerings in Moche burials (Goepfert 2012), suggesting a significant role for the carcasses of these animals in ritual activity associated with the dead and perhaps for the procurement of these animals as well. Nevertheless, domestic animals (dogs, guinea pigs, camelids) are far more prominent in sacrificial and ritual contexts, such as in association with human burials or building construction (Goepfert 2012; Sandweiss and Wing 1997). The primacy of domestic animals as the appropriate substrate for sacrifice is supported in the ethnohistoric literature. Cobo (1990 [1653]:113) recorded that for an object (animal or otherwise) to be appropriate for sacrifice, it must be one that required human labor to produce. While this line of thinking may be largely utilitarian, it aligns well with the notion that the lived interaction between humans and animals is significant within the context of ritual sacrifice. A wealth of archaeological, ethnohistoric, and ethnographic information documents the ritual importance of camelids in the Peruvian Andes and highlights ceremonies in which camelids were sacrificed (for a comprehensive review, see Bonavia 2008). Sources describe the importance of camelid sacrifice in Inka ritual activity, with the ritual killing of camelids being a regular event in herds controlled either by local communities or the religious arm of the Inka state (often referred to as “church herds” or “temple herds”) and the Inka royalty (Flannery et al. 1989; Murra 1965). The scale of these sacrifices ranged considerably from single animals to several hundred individuals of very specific types (e.g., specifically colored camelids that were sacrificed at certain times of the year or were associated with particular deities) (Murra 1965). On the north coast of Peru, camelid skeletal elements are relatively common items in Moche (AD 100–800) graves, with the most common type of faunal funerary offering consisting of camelid skulls and distal limb elements, typically from all four limbs (Donnan 1995; Goepfert 2012; Gumerman 1994; Millaire 2002). Such remains do not provide direct evidence of ritual killing, and it has been suggested that the bulk of the animals were consumed during funerary feasting (Goepfert 2010). Camelid remains have also been found in association with ritually killed humans on the north coast of Peru, such as at Huaca Santa

3 2 6   Szpak et al.

Clara in the Virú valley (AD 1100; Millaire 2015; Szpak et al. 2014), Cerro Cerrillos in the Lambayeque valley (AD 900–1100; Klaus et al. 2010), Santa Rita B in the Chao valley (AD 1100–1300) (Gaither et al. 2008), at Túcume in the La Leche valley (Toyne 2011b), Huanchaco in the Moche valley (AD 1400; Donnan and Foote 1978), and the ChotunaChornancap Archaeological Complex in the Lambayeque valley (AD 1450–1532; Wester et al. 2010). In some cases, there is clear evidence that these camelids were sacrificed in association with human ritual killing, such as at Huanchaco, where the camelids show evidence of blunt force trauma to the cranium (Donnan and Foote 1978). In other cases, the camelids do not display clear physical evidence of unnatural death, but the deposition of entire, unbutchered camelid carcasses contemporaneous with the human ritual killings is strongly suggestive of unnatural deaths. Outside contexts of human ritual killing, the sacrifice of camelids specifically is well documented on the north coast of Peru during the late Middle Horizon and beginning of the Late Intermediate Period. The camelid remains analyzed in this study are associated with one such ritual event at Huancaco in the lower Virú valley and is the earliest documented instance of a series of camelid sacrifices (which in some cases included human subadults as well) on the north coast of Peru. In the Virú valley, similar sacrifices occurred at Huaca Santa Clara (V−67), Huaca Negra (V−71), and Cerro de Huarpe (V−313) (Millaire 2015; Strong and Evans 1952; Willey 1953). The number of camelids sacrificed at each of these sites varied from four (V−313) to twenty-eight (V−67), but all were associated with previous occupation centers that had been abandoned for some time prior to the sacrificial event. The association of these camelid sacrifices with previously occupied sites strongly suggests the significance of ancestral places on the landscape, although these sites did not function as ceremonial centers during their occupations (Millaire 2015). On the north coast of Peru and throughout the central Andean region, the variable nature of camelid offerings that occur on their own or in association with human burials (or ritual killings) makes it difficult to reconstruct any generalized patterns. This may be because camelids were utilized in a diversity of ritual contexts, being sacrificed in different manners (mode of death, color of animal, age of animal, number of animals) depending on the local context. Even within the north coast camelid sacrifices described above there is considerable variation. This

Life Histories of Sacrificed Camelids  327

is not surprising in light of the key roles that camelids played in ritual activities as documented in historic and modern times (Bolin 1998; Flannery et al. 1989; Flores-Ochoa 1979; Kolata 1996; Murra 1965; Webster 1972; Zuidema 1992). This is not to suggest that there is direct continuity between any particular camelid ritual over the last millennium, but given the ubiquity of camelids in the region, it is likely that they have played an important role in rituals over that time.

Stable Isotope Variability in the Andes Stable isotope analysis has been employed with increasing regularity in archaeological studies to reconstruct the life histories of past human and animal populations, including in the Andean region of South America (Finucane et al. 2006; Finucane 2007; Kellner and Schoeninger 2008; Knudson et al. 2007; Slovak and Paytan 2011; Williams and Katzenberg 2012). In the case of human and animal organic tissues, the isotope systems studied can be broadly characterized as either dietary (as in the case of C, N, S) or geographic (as in the case of O, H, Sr, Pb), depending on the biogeochemical processes that primarily impact a particular isotope. Ultimately, however, the dietary isotope ratios will also be influenced by geography because various environmental processes at global, continental, and regional scales can influence the carbon, nitrogen, or sulphur isotopic compositions at the base of the food web (Craine et al. 2009; Murphy and Bowman 2006; Szpak et al. 2013; Tieszen and Chapman 1992). The foci of this chapter are the stable isotopes of carbon and nitrogen, which are routinely analyzed to examine the diet and ecology of past human and animal populations. Because the carbon and nitrogen isotopic compositions of an animal’s tissues reflect the carbon and nitrogen isotopic compositions of the foods consumed (DeNiro and Epstein 1978, 1981), it is first necessary to review some of the processes that influence these signatures and what might be expected in northern Peru. In terrestrial ecosystems, the majority of plants consumed by animals in appreciable quantities are of two photosynthetic types: C4 plants (mean δ 13C ca. −12 ‰), which are tropical grasses (including maize and amaranth); and C3 plants, which comprise most other terrestrial vascular plants (including some grasses, mean δ 13C ca. −26 ‰) (Kohn 2010; O’Leary 1981; Smith and Epstein 1971). Relative to C3 plants, C4 plants tend to be more abundant in hot and dry locations (Sage and Pearcy 2004). Along the western slope of the Andes, mean daily temperature

3 2 8   Szpak et al.

tends to decrease with altitude, and mean annual precipitation tends to increase with altitude (Bush et al. 2005). Therefore, C4 plants are expected to be more abundant at lower altitudes relative to higher altitudes, and this pattern has been observed in several studies (Szpak et al. 2013; Tieszen and Chapman 1992). The regional significance of altitude in influencing the isotopic compositions of plants extends beyond the relative distribution of C3 and C4 species. The extremely arid conditions characteristic of most lowaltitude and coastal locations in northern Peru result in a trend toward relatively high δ 15N values in plant tissues relative to higher-altitude locations that are relatively more moist (Szpak et al. 2013). This effect has been observed throughout the world and is driven by the fact that hot and arid ecosystems tend to be fairly “open” and prone to various processes that lead to the loss of nitrogen (these processes favor the loss of 14N) and, concurrently, the relative enrichment in 15N (Handley et al. 1999). Significantly, these effects are passed on to herbivores (Murphy and Bowman 2006). In northern Peru, these aridity effects may be dampened or absent for agricultural plants relative to wild plants growing on the coast due to the supplementation of agricultural plants with irrigation water (Szpak et al. 2012). Overall, there is a trend toward higher carbon and nitrogen isotopic compositions in plants growing in low-altitude and coastal locations relative to high-altitude locations. In addition, there is a trend toward increasing baseline isotopic variability at coastal and low-altitude locations (Szpak et al. 2013). Besides the potential effects of irrigation water discussed above, there are also significant numbers of nitrogen-fixing plants known to have grown extensively along the coast, for instance: peanuts, various types of beans, and a number of different pod-bearing Prosopis trees (Beresford-Jones et al. 2009; Pozorski 1979). This is significant because nitrogen-fixing plants are characterized by lower δ 15N values relative to plants that obtain their nitrogen from mineralized nitrogen (nitrate and ammonium) in the soil (Virginia and Delwiche 1982). Thus, coastal vegetation is extremely diverse in terms of carbon and nitrogen isotopic compositions, with all of the following being present (or at least plausible): low δ 13C/low δ 15N (nitrogen-fixing plants, C3 cultigens grown under irrigation), low δ 13C/high δ 15N (wild C3 plants, fertilized C3 plants), high δ 13C/low δ 15N (cultivated maize grown under irrigation), and high δ 13C/high δ 15N (wild C4 plants, fertilized maize) (Szpak et al. 2013).

Life Histories of Sacrificed Camelids  32 9

Archaeological Context The site of Huancaco (V−88 and V−89; Willey 1953) is located in the lower Virú valley, approximately 8km from the coast and 5km east of the Gallinazo Group (figure 12.1). Initially, this site was believed to have been occupied during the Early Intermediate Period, functioning as the paramount Moche regional administrative center of Virú (Fogel 1993; Willey 1953). Excavations at Huancaco by Bourget (2003, 2010) document the occupation between AD 550–6801 and demonstrate an absence of evidence for the presence of Moche administrators. Instead, the site was occupied by a group of elites characterized by a distinct but related style of material culture (Huancaco)—one of several regional material culture traditions on the north coast (Chapdelaine 2008; Donnan 2009; Millaire 2009a). Fourteen of the camelids analyzed are associated with the Early Intermediate period occupation of the site. Ten of the camelid remains analyzed in this study postdate the abandonment of Huancaco and are associated with a ritual that involved the

12.1. Images of select camelids from the post-abandonment ritual sacrifice at Huancaco. Photos by Steve Bourget.

3 3 0  Szpak et al.

sacrifice of 15 immature llamas atop the abandoned site (figure 12.1).2 Following the abandonment of the site (ca. AD 700), there is no evidence for its reoccupation aside from the material associated with this ritual. The llamas were recovered from shallow pits dug into two rooms associated with the earlier occupation. The ritual appears to have occurred sometime between AD 778–979 on the basis of a radiocarbon date on a cord used to bind one of the llama’s legs (Millaire 2015). Such reuse of abandoned huacas and deposition of ritual offerings occurred regularly on the north coast of Peru. On the basis of skeletal and dental development (Wheeler 1982), the age range of the llamas was very restricted: ten animals were between 1 and 3 months old, four were between 5 and 6 months old, and one was between 6 and 9 months old. While a specific age range of the llama population was targeted, there was considerable variation in the treatment of the llamas in this ritual and their subsequent burial. Of the 15 llamas, nine were buried individually and six were buried in pairs. Some llamas, but not all, had their feet bound. Of the three paired llama burials, a younger individual (1–3 months) was found below an older individual (4–6 months). This may be coincidental or may represent a preference to position the animals in this way. Alternatively, this may be caused by a punctuated temporal sequence, with burials taking place over a longer period of time. There is, however, no evidence to support this latter possibility. The stomachs and esophagi of all of the undisturbed animals recovered contained trapezoidal pieces of Spondylus shell and rectangular pieces of turquoise (figure 12.2). There are no cut marks on the ribs suggesting that these objects were force-fed to these animals immediately prior to their deaths. While it is possible that force-feeding these objects caused the animals to asphyxiate, there is no definitive evidence to this effect.

Materials and Methods Bone collagen was extracted using a modified Longin (1971) method as described in detail elsewhere (see Szpak 2013). Hair samples with adhering skin were manually cleaned and sampled in 1cm increments. Hair was cleaned by sonication in deionized water and subsequently treated with 2:1 chloroform: methanol. Hair samples were air-dried and finely chopped prior to isotopic analysis. Isotopic and elemental compositions were determined using either

Life Histories of Sacrificed Camelids  331

12.2. Detailed image of Llama 2 with turquoise and Spondylus objects force-fed to the animal at or around the time of death. Photo by Steve Bourget.

a Thermo Finnigan Delta V or a Thermo Finnigan Delta Plus XL continuous flow mass spectrometer coupled to a Costech Elemental Analyzer at the Laboratory for Stable Isotope Science at the University of Western Ontario. Samples analyzed on the Delta V were calibrated to VPDB and AIR with two glutamic acid standards: USGS40 (accepted values δ 13C = −26.39‰, δ 15N = −4.5‰) and USGS41 (accepted values δ 13C = +37.63‰, δδ 15N = +47.6‰). Samples analyzed on the Delta Plus XL were calibrated to VPDB with IAEA-CH-6 (sucrose, accepted δ 13C = −10.449‰) and NBS-22 (oil, accepted δ 13C = −30.031‰) and to AIR with IAEA-N-1 (ammonium sulfate, accepted δ 15N = +0.4‰) and IAEA-N-2 (accepted δ 15N = +20.3‰). In addition to these calibration standards, internal (Keratin, average δ 13C = −24.04‰, δ 15N = +6.4‰) and international (IAEA-CH-6 and IAEA-N2 where USGS40 and USGS41 were used as calibration standards) standards were analyzed to monitor analytical accuracy and precision. Ten percent of samples were analyzed in duplicate with a mean difference among five duplicate pairs of 0.04‰ for δ 13C and 0.03‰ for δ 15N.

Results Carbon and nitrogen isotopic compositions for camelid bone collagen are presented in table 12.1 and figure 12.3, along with modern animals

3 3 2   Szpak et al. Table 12.1. Carbon and Nitrogen Isotopic Data for Huancaco Camelids Sample ID

% Collagen

C:N Ratio

δ13C (‰, VPDB)

δ15N (‰, AIR)

Sacrificed llamas Llama 1

2.8

3.38

−14.50

+6.5

Llama 4

4.1

3.11

−17.08

+10.1

Llama 5

1.3

3.19

−18.55

+6.7

Llama 6

7.4

3.02

−17.10

+7.2

Llama 7

7.7

2.99

−13.97

+8.6

Llama 8

3.9

3.04

−16.42

+8.7

Llama 9

10.9

3.00

−16.15

+7.2

Llama 10

16.9

3.22

−15.99

+6.7

Llama 11

5.8

3.22

−15.71

+6.5

Llama 13

3.6

3.26

−16.75

+4.9

2.93

−15.00

+6.6

Occupation-associated camelids AIS 1214

22.1

AIS 1221

20.3

2.96

−18.46

+6.6

AIS 1222

17.3

2.95

−14.14

+7.9 +7.2

AIS 2131

12.7

3.14

−12.00

AIS 2133

7.2

3.14

−13.15

+7.0

AIS 2134

15.0

3.15

−12.05

+8.5

AIS 2136

16.8

3.16

−15.81

+5.0

AIS 2137

1.9

3.37

−18.11

+10.7

AIS 2138

4.1

3.14

−10.83

+5.7

AIS 2140

16.8

3.16

−15.98

+7.3

AIS 2142

15.0

3.17

−14.76

+6.1

AIS 2147

10.3

3.16

−10.64

+4.6

AIS 2148

5.7

3.18

−11.64

+6.7

AIS 2149

8.8

3.22

−15.48

+6.1

Data from Llama 2 not included as it is considered unreliable (see text).

raised in the highlands for comparison. For bone collagen, only samples that met the following criteria were included in subsequent analyses: (1) collagen yield >1%; (2) C:N ratio between 2.9 and 3.6; and (3) %C > 13%, %N > 4.8% (Ambrose 1990; DeNiro 1985). The carbon isotopic compositions for the sacrificed camelid bone collagen (n = 10) ranged from −18.55 to −13.97‰; nitrogen isotopic com-

Life Histories of Sacrificed Camelids  333

positions ranged from +4.9 to +10.1‰; there was no correlation between δ 13C and δ 15N. The carbon isotopic compositions for the Early Intermediate Period occupation camelid bone collagen (n = 14) ranged from −18.64 to −10.64‰; nitrogen isotopic compositions ranged from +4.6 to +10.7‰. While the turnover rates of bone collagen in camelids are not known specifically, the isotopic signature of bone collagen represents the weighted average of dietary intake over the last years (or decades) of an animal’s life (Wild et al. 2000); bone formed early in life during periods of rapid growth may be disproportionately represented (Hedges et al. 2007). Because the sacrificed llamas analyzed in this study are all less than 9 months old, it is reasonable to simply state that the isotopic composition of their bone collagen represents lifetime average diet. Carbon and nitrogen isotopic compositions for serially sampled hair from the sacrificed llamas are presented in figure 12.4. For serially

12.3. Carbon and nitrogen isotopic compositions of Huancaco camelid bone collagen. Modern camelid bone collagen data (δ 13C adjusted by +1.5‰; Yakir 2011) from high-altitude contexts are shown for comparative purposes (Schoeninger and DeNiro 1984; Szpak 2013; Thornton et al. 2011). Image by Paul Szpak.

3 3 4   Szpak et al.

12.4. Carbon and nitrogen isotopic compositions of serially sampled hair from Huancaco llama sacrifices. Numbers to the left of each segment line correspond to the llama specimen. Image by Paul Szpak.

sampled camelid hair, there were no consistent patterns leading up to the time of death, with some individuals (e.g., Llama 5) having very low variation over time (range of 1‰ for δ 13C, 0.6‰ for δ 15N), and with other individuals (e.g., Llama 4) having much larger variation over time (a range of 0.9‰ for δ 13C, 4.3‰ for δ 15N). Moreover, there is no consistent direction in the dietary change leading up to the time of death, with individuals variably having increasing, decreasing, or largely stable δ 13C or δ 15N values. The sacrificed Huancaco camelids had significantly lower δ 13C values relative to the Early Intermediate Period non-sacrificed camelids (t-test, t[22] = 2.37, p = 0.03), although there was considerable overlap between the two groups. The δ 15N values did not significantly differ for the sacrificed and Early Intermediate period camelids (t-test, t[22] = 0.73, p = 0.47). Although the sacrificed juveniles were characterized by higher δ 15N values than the occupational camelids (most of which were mature), the difference was small (+7.3 ± 1.5‰ compared to +6.9 ± 1.5‰). This is

Life Histories of Sacrificed Camelids  335

surprising, considering the restricted age range of these camelids and the potential impact of nursing on δ 15N values. Today, most llamas are weaned by 6 months of age (Fowler 1998). On this basis, we would expect that these animals present a fairly strong “suckling effect,” causing higher δ 15N values than would be expected in adult animals (Balasse et al. 2001). By way of comparison, a larger group of immature sacrificed camelids from Huaca Santa Clara in the Virú valley also had δ 15N values that were only marginally higher than a group of butchered animals (primarily mature) from the same site (+7.2 ± 1.3‰ compared to +6.7 ± 1.4‰) (Szpak et al. 2014). That there is not a more pronounced suckling effect may be related to the very young age of these animals, most of which were less than 3 months of age at death. Extremely young animals tend to have tissue nitrogen isotopic compositions that are similar to the mother, acquiring the isotopic suckling effect after a prolonged period of nursing (Fogel et al. 1989; Schurr 1997). Therefore, these animals may not have lived long enough to record a substantial nursing signal in the nitrogen isotopic composition of their bone collagen.

Discussion Geographic Origin of Camelids Carbon and nitrogen isotopic compositions of animal tissues are, by themselves, not sufficient to definitely determine the local nature of an individual or group of individuals. As discussed previously, on the coast there are more plants with high δ 13C values (C4 plants) and high δ 15N values (due to the arid conditions). These plants, however, do not make up the entirety of the local plant biomass, and numerous species with low δ 13C and δ 15N values are common; therefore, animals with low δ 13C and δ 15N values are not necessarily foreign. Conversely, plants with high δ 13C and high δ 15N values are rare or absent at high altitudes typically associated with camelid pastoralism (Szpak et al. 2013). This is evident when isotopic data from modern camelids living in high-altitude locations are considered (figure 12.3). These data are characterized by low δ 13C and δ 15N values, with no individual δ 13Ccollagen > −17.7‰ (after adjustment for the Suess Effect; Yakir 2011). Similarly, archaeological data for camelids from high-altitude sites are consistent with a strong or complete reliance on C3 plants (DeNiro 1988; Thornton et al. 2011). It is therefore reasonable to suggest that animals with high δ 13C and δ 15N values are not likely to have been raised in the high sierra, or puna.

3 3 6   Szpak et al.

Along these lines, it is likely that at least some of the camelids involved in this sacrificial event were raised locally on the coast. It is possible that those with low δ 13C and δ 15N values were imported from elsewhere for the explicit purpose of this sacrificial event, but this scenario is unlikely given the restricted age range of these camelids. Based on modern and historic accounts of long distance caravan travel by camelids in the Andes, the caravan animals were exclusively castrated males older than two years of age (Browman 1990; Nielsen 2001); there are no records of juvenile camelids making these long treks from the highlands to the coast. The most reasonable explanation for the presence of juvenile animals is that these animals were drawn from a relatively local population. If these sacrificed camelids were adults, this assertion would be much more problematic. This is certainly not to imply, however, that the local inhabitants maintained large camelid herds that were quantitatively or qualitatively similar to what has been described in the highlands (Flores-Ochoa 1979; McCorkle 1987; Murra 1965, 1980; Orlove 1977; Tomka 1992). Rather, a more likely scenario would see a smaller scale camelid pastoralism along the coast and in the lower stretches of the coastal river valleys that were largely subsidized by agricultural and marine food sources (Szpak et al. 2014). Several authors have proposed the herding of camelids outside of the highlands on the basis of isotopic information (DeNiro 1988; Goepfert et al. 2013; Thornton. 2011) and other lines of evidence (Goepfert 2012; Pozorski 1976; Shimada and Shimada 1981, 1985). While the isotopic data in and of themselves do not solidify the origin of these animals as local, the age structure of this group of animals is strongly suggestive of this possibility. Life Histories of Sacrificed Camelids Although these animals were most likely born at or around the same time (probably within 6 months of one another) and eventually killed at this location (probably simultaneously), there is no apparent consistency in their diets based on bone collagen and hair isotopic data. This may be because these animals were either from different locations or were managed in different ways. Given the existing isotopic data, it is not possible to definitely suggest one as more likely. Irrespective of geographic origin, however, the isotopic data do indicate that there was some variation in the management of these sacrificial animals. If these

Life Histories of Sacrificed Camelids  3 37

sacrifices represented the culling of individuals from a single large herd, we would expect to see more consistency in the isotopic variation in camelid hair sampled over time (similar diets leading up to the time of death reflected in the hair), as well as less isotopic variation among individuals (similar diets averaged over the life of the animal reflected in the bone collagen). While there is less variation in carbon isotopic composition in the sacrificed camelids relative to the Early Intermediate period camelids (figure 12.3), this may be driven by the smaller sample size of the sacrificed camelids, their extremely restricted age range, and/or their restricted temporal range relative to the Early Intermediate Period camelids. There is no indication that these sacrificed animals spent any appreciable time together prior to being killed as part of this ritual. If these animals were part of a single herd and were grazing together prior to the sacrificial event, then a pattern of more limited isotopic variation (comparable to that observed for the modern highland camelids) would be expected. The similarity of isotopic compositions between the occupation and sacrificial camelids is significant when considering the selection of these specific camelids to be sacrificed. Common themes in animal sacrifice include the surrogacy of individual animals for individual people, as well as the importance of some degree of ownership of the animal being sacrificed. In this case, the isotopic variation that existed among individual camelids sacrificed at this site may very well be driven by differences in animal management by specific herders. The inclusion of animals derived from different herds (and possibly distinct locations) may reflect the contributions of distinct people or socioeconomic collectives to this ritual event. The possibility exists that these animals were acquired from numerous herders prior to the sacrifice (through exchange or some form of tribute) and that the sponsor of the sacrificial event was divorced from the life histories of the camelids. In other words, the people involved in the rearing of the animals appear removed from the sacrificial event, and the selection of these particular individuals may be incidental. Instead, the driving force behind the sacrificial event, in the broadest sense, may have been the selection of a particular location to bury these animal offerings. The subjectivity of the camelids was largely removed, with their primary purpose being for visual effect during the performance of the ritual or for serving as materializations of the spiritual purpose of the event.

3 3 8   Szpak et al.

Camelid Sacrifice in Broader Context Goepfert (2012) has suggested that the pervasiveness of immature camelids in sacrificial contexts on the north coast is evidence of the existence of large local herds. A large size was required to buffer against the effects of culling these animals for sacrifice. While this is a possibility, evidence for the existence of large corrals on the coast is lacking, although some smaller corrals have been noted (Bawden 1982; Shimada 1994a; Toyne 2008; Wilson 1988). The lack of corrals on the coast is significant given the lack of large pastures on the coast that are comparable in size to those in the puna; the large agricultural fields in coastal river valleys would certainly not have fared well with large camelid herds grazing freely. An alternative scenario—large numbers of small camelid herds distributed throughout coastal river valleys—seems more plausible given the nature of the coastal environment and is more consistent with the isotopic data (Szpak et al. 2014). In this context, a relatively large local population of camelids could be maintained, although their management would occur at a much smaller scale, with small numbers of individuals being kept by families or other small social units (Szpak et al. 2014). This smaller-scale management of camelids is significant in terms of the inclusion of animals for sacrifice. As discussed above, the isotopic differences observed among individuals, as well as the varying patterns in isotopic compositions within individuals (hair), may very well represent the involvement of numerous people (or groups of people) donating camelids for this ritual. The structuring of the sacrifice in this respect may have been necessitated by the relatively small size of herds outside of prime camelid pastureland, with no single herd being large enough to sustain the sacrifice of more than a dozen (or in some cases, dozens) individuals in a given season. Aside from these practical considerations, the ritual event itself may have been significant as a unifying mechanism for the larger community, potentially to honor and communicate with ancestors, to articulate with ancestral places on the landscape, and as a spectacle (Hoskins 1993). The isotopic variation in these sacrificial camelids may very well represent the bringing together of animals raised under different circumstances by different people; it is not unusual in cases of animal sacrifice for particular animals to be closely associated with the people who raised them, with some degree of animal-person substitution involved (Küchler 2002).

Life Histories of Sacrificed Camelids  339

That the animals sacrificed were juveniles is undoubtedly significant. Llamas may have been utilized for their meat, but the value of their secondary products was extremely important. These products most notably include fiber for textile production (Splitstoser 2009) and dung for use as fertilizer and fuel (Szpak et al. 2012; Winterhalder et al. 1974). These animals were sacrificed prior to being shorn, and they were not consumed. Considering sacrifice in terms of a gift economy with the supernatural, this may signify a more extravagant offering given to the ancestors relative to an aged, infertile animal (Hoskins 1993) or body parts with low meat value—a common occurrence in Moche funerary contexts (Goepfert 2012). It is important to bear in mind that sacrifices are not events confined to the ritual sphere. In discussing Old World camel sacrifice, Rahimi (2004: 452) points out that “the performance of the rituals draws attention to an inventive socio-religious hybridization process between practices of heterodoxy and canonical-clerical interpretations, forming a ritual space that essentially overlaps with the political sphere.” This would certainly have been true of these camelid sacrifices on the north coast of Peru, which emerged after the collapse of the large Early Intermediate Period polities (possibly the earliest states in the region) at a time when there was receptiveness to new cultural traditions, both internal and external (Millaire 2015). One of the key elements of these north coast camelid sacrifices is that the isotopic data suggest continuity in the manner of animal husbandry for the llamas that were sacrificed versus those that were raised and butchered during the occupation of these sites (Szpak et al. 2014), which may well represent the melding of local camelid husbandry traditions with new ritual practices.

Conclusion The llama sacrifices at Huancaco represent one ritual event in a larger regional tradition of sacrifice involving llamas and, in some cases, human children on the north coast of Peru. The carbon and nitrogen isotopic data suggest that the animals used in the ritual were drawn from a fairly local range, probably from within the valley. The isotopic variation observed within the sacrificed group of camelids is inconsistent with these animals consuming similar diets for any significant amount of time, suggesting that they were managed separately prior to the ritual event. Additional studies of this nature on other sacrificed camelid

3 4 0  Szpak et al.

groups on the north coast would be useful in documenting regional variability in this practice. Additionally, where humans and camelids occur together in these contexts, there may be some interesting geographic connections between people and animals. Were the humans involved in the ritual drawn from a similarly local area? Are the isotopic patterns similar for rituals involving only camelids versus those involving both camelids and humans? Are there any geographic connections between humans and camelids that are directly associated with one another (i.e., those buried in the same pit)? With respect to animal sacrifice, if our goal is to understand these events within a broader social and cultural context, then it is important to better understand the lives of these animals in addition to reconstructing the circumstances surrounding their deaths. Stable isotope analysis offers one means to better reconstruct the life histories of animals, both at the individual and supra-individual levels that ultimately speak to a previously unknown range, process, and structure surrounding animal offerings on the north coast of Peru. There is a suite of other techniques, besides those described here, that would aid in the reconstruction of the lives of these animals. Additional isotopic markers (δ 18O, δ 2H, 87Sr/86Sr) that more specifically address habitat and mobility, rather than diet, have the potential to document the geographic origin of animals that were part of ritual activities, including sacrifice. For example, Knudson et al. (2012) used strontium and oxygen isotope analyses to examine the geographic origin of camelids consumed during feasting at Tiwanaku. Such studies are dependent upon knowledge of local isotopic baselines, as well as sufficient isotopic variation among regions (i.e., two locations may be geographically isolated from one another but may share similar oxygen, hydrogen, or strontium isotope signatures). While not in sacrificial contexts, the examination of pathological lesions on camelid bones has documented the presence of cargo animals in southern Peru (deFrance 2010) and northern Chile (Labarca and Gallardo 2012). The development and advancement of next-generation DNA sequencing technologies make it possible to attain genomic-level perspectives of ancient organisms (Knapp and Hofreiter 2010). Although methodologically difficult at this time, it is theoretically possible to use these data to reconstruct various phenotypic traits, including externally visible characteristics (Fortes et al. 2013). These analyses could allow various aspects of an animal’s appearance to be reconstructed, as well as identify genetic relationships to other

Life Histories of Sacrificed Camelids  341

animals involved in ritual events. Taken together, these techniques allow for a much more detailed understanding of the lives of animals involved in ritual killings. When interpreted alongside detailed contextual information, these lines of evidence have the potential to reveal a great deal regarding the ritual killing of animals on the ancient north coast of Peru. Acknowledgments Estuardo La Torre coordinated sample collection, and Steve Bourget provided access to the samples. Kim Law and Li Huang provided technical assistance. Sharon Buck, Rebecca Dillon, and Rebecca Parry assisted with sample preparation. This project was supported by the Wenner-Gren Foundation (Dissertation Fieldwork Grant #8333 to P.S.), Social Sciences and Humanities Research Council of Canada (Standard Research Grant to C. D. W., F. J. L., and J-F. M.), Natural Sciences and Engineering Research Council Discovery Grant (F. J. L.), Canada Foundation for Innovation and Ontario Research Fund Infrastructure grants (F. J. L.), Canada Research Chairs Program (C. D. W. and F. J. L.), and the University of Western Ontario. Isotopic research was conducted under Resolución Viceministerial No. 014-2013-VMPCIC-MC. The editors of this volume and three anonymous reviewers provided editorial reviews of earlier drafts of this chapter.

Notes 1. All dates are calibrated using IntCal09 (Reimer et al. 2009). 2. While llamas and alpacas are notoriously difficult to distinguish on the basis of postcranial skeletal remains, differentiation between the species is possible on the basis of dentition and when soft tissue is present (Wheeler 1982; Wheeler, et al. 1995). In the case of Huancaco, butchered animals from the Early Intermediate Period occupation of the site cannot be identified to species and are simply referred to as “camelids.” Conversely, sacrificed animals from the later ritual context are complete, have retained soft tissue, and can be classified as llamas specifically.

3 4 2   Millaire

chapter thirteen

Posts and Pots Propitiatory Ritual at Huaca Santa Clara in the Virú Valley, Peru Jean-François Millaire Department of Anthropology, The University of Western Ontario, London, ON, Canada

In selecting a title for this volume of essays, the editors explicitly placed the concept of ritual killing at the center of sacrificial practice in the ancient Andes. In so doing, many chapters force us to consider more closely ritual actions that involved one form or another of violence toward human beings. This is a valuable exercise, but it is one we should embark on with caution, resisting the temptation to restrict the concept to bloodied corpses and recognizing that nonliving objects, too, are often killed or terminated as part of sacrifice ritual through disfiguring, breakage, destruction, or total obliteration. Although human sacrifices clearly trigger something visceral in all of us, I argue here that we should probably be wary when elevating human ritual killing to a special status detached from the wider set of sacrifices performed by individuals, past or present, on a more regular basis. Here I have no intention to discuss human sacrifices sensu stricto but wish to consider sacrifice more broadly: that is, as a fundamental act of worship and a keystone of relations with the supernatural in premodern civilizations in which a range of items were offered to divinities in order to establish, maintain, or restore a right relationship between humans and the sacred order (Faherty 2011). From this perspective, human blood was often considered the most valuable item many Andean societies

Posts and Pots  3 4 3

agreed to part with in order to transfer life-force or power to the supernatural or to fulfill the ideological or political goals of those in control. Yet many other types of materials also had symbolic currency. The rituals discussed in this chapter took place during the sixth century AD inside a large room best described as an open gallery on a Virú settlement known today as Huaca Santa Clara. This room’s founding and construction began with a propitiatory ritual, the precise meaning of which is still difficult to grasp. This ritual is nevertheless interesting in that it reminds us of the complexity of the processes at play behind every and any act of sacrifice (from simple fasting to the execution of prisoners as part of a mass killing), and of the fact that much hinges on what is deemed a satisfactory relinquishment at a specific or situational junction in time and space rather than simply on the substance of what is being surrendered.

Huaca Santa Clara Huaca Santa Clara was an agglutinated settlement perched on the flanks of a hill (Cerro Cementerio) that dominates the lower Virú valley, a river oasis located south of present-day Trujillo on the north coast of Peru (figure 13.1). The settlement functioned as an administrative center that was part of the valley-wide Virú state system1 during the Early Intermediate Period (Larco Hoyle 1945; Millaire 2009a, 2010a, 2010b; Willey 1953). This state formation featured a valley-wide irrigation system as well as a four-tiered settlement network with a capital city that was located at the Gallinazo Group in the lower valley (Bennett 1950; Millaire and Eastaugh 2011), administrative and defensive settlements (e.g., Huaca Santa Clara), villages, and hamlets. In 2002 and 2003, work was carried out at Huaca Santa Clara, documenting the Virú occupation of the settlement (Millaire 2010a) and its late reoccupation for a spectacular ritual event (Millaire 2015; Millaire and Surette 2011). Radiocarbon dates obtained from this site suggest that the occupation could date back to the first and possibly second century BC and that it lasted until the seventh or eighth century AD (table 12.1). All dates were obtained by radiometric analysis, and calibrations were generated using the IntCal09 calibration curve (Reimer, et al. 2009) in OxCal 4.1. This ancient settlement covers the entire surface of Cerro Cementerio (figure 13.1, figure 13.2), a triangular-base hill that features a civic

3 4 4   Millaire

13.1. The site Huaca Santa Clara on Cerro Cementario in the distance. Photo by Jean-Françoise Millaire.

building on its apex, three adobe brick platform terraces at mid-height on each edge, and residential areas in the lower parts. The site’s administrative character is evidenced by the overall design of the architectural complex, by the complexity of the access patterns, by the nature of the portable objects uncovered, and, most important, by the presence of a system of agglutinated storage rooms on each of the three mid-height terraces (Millaire 2010a). The site was originally dominated by a civic building largely destroyed by modern activities. It consisted of a stepped adobe platform adorned with decorative war maces made of baked clay, a type of architectural ornament well known in north coast art. The platform overlooked a system of storage rooms perched on the mid-height terraces (20–30m above the valley floor). About a hundred small rectangular chambers

Posts and Pots  3 4 5

13.2. Aerial photograph of Huaca Santa Clara with its architecture highlighted. Image by Jean-François Millaire.

organized in honeycombed patterns were identified. These rooms featured walls up to 2m high and were originally accessed through the roof, possibly using ladders. Although most rooms are now empty, the presence of large quantities of organic materials in some spaces suggests that they were originally used for the storage of food crops from the surrounding agricultural lands.2 This evidence led us to hypothesize that Huaca Santa Clara was directly involved in the storage of produce harvested from the immediate area and was possibly part of some form of state taxation and redistribution system (Masur et al. submitted; Millaire 2010a, 2010b). The mid-height terraces also featured spaces for public gatherings, as well as residential structures presumably used by members of the local elite, as indicated by the access pattern and quality of the architecture.

3 4 6   Millaire

Life and Death at Huaca Santa Clara One of the most interesting spaces uncovered at Huaca Santa Clara was a large open gallery (3m x 10.5m) located on the eastern side of the hill (Room A-102), one of the mid-height terraces (figure 13.3). It was part of a large multiroom building (40m x 20 m) that featured a stepped façade to the south. Access to this area was possible through a baffled entrance located on the eastern side of the platform, which led visitors to two adjacent rooms. The second room (a raised platform) featured a wall with niches and a low bench for a single person to sit. A narrow passage provided access to even more private areas, including the open gallery that

13.3. Room A-102 at Huaca Santa Clara. Photo by Jean-François Millaire.

Posts and Pots  3 47

13.4. Isometric reconstruction of Room A-102. Drawing by Jean-François Millaire.

interests us here (figure 13.4). This room was architecturally exceptional in that it featured three large wooden posts (30cm in diameter) that originally held transverse beams supporting a slanting roof and a low eastern wall that formed a balustrade. This open design offered a panoramic view to the eastern flank of the settlement, the adjacent agricultural lands, and the valley neck—a narrow pass through which precious water and chaupiyunga (mid-valley) and highland commodities reached the coast. This gallery clearly set its occupants apart from the local populace, although different lines of analysis revealed that local elites probably did not live the ostentatious life typical of state leaders in the Gallinazo Group capital (Bennett 1950; Millaire 2010b). Indeed, although Huaca Santa Clara leaders had access to exotic camelid wool to weave brightly decorated clothing (Millaire 2009b), the architecture remained modest, and non-utilitarian ceramic vessels were in short supply, especially those pots that were obtained through long-distance trade or diplomatic relations with neighboring polities (Millaire 2010a). Besides its potential function as a lookout, this open gallery would have been ideal as a living space to greet and entertain people, eat, sleep, or craft objects (no evidence of food preparation was found). Based on work carried out on other sites in this region, we also know that household-level rituals were extremely common and included libations, burning of offerings, and ritual killing of animals, often as part of the daily preparation of meals (Cutright 2009; Tsai 2012). That being said, traces of such rituals are only rarely preserved in the archaeological record, and none survived inside this gallery.

3 4 8   Millaire

13.5. Burial 7, Room A-102. Photo by Jean-François Millaire.

The floor of the room was found approximately 1m below the existing surface beneath layers of refuse and broken adobes, suggesting that the space had been intentionally filled to serve as a foundation for a subsequent building, a common construction technique at the site. Prior to being filled, however, the room was the theater of a funerary ceremony, as indicated by the presence of two burials along the west wall. The corpses had been placed inside a cut and subsequently covered with a mix of ash and sand, which, together with the naturally dry climate on the slopes of the hill, contributed to the excellent preservation of the remains (Dillon 2013). The first individual was an adult female (20–35 years old) on her back with her head oriented toward the north (Burial 7; figure 13.5). Her body was in an excellent state of preservation. The corpse had been naturally mummified, and tattoos were still visible on her face. She was wearing a textile tunic, and her body was rolled in three successive shrouds. She was buried with a cotton pouch, a single gourd bowl, and a dog—a relatively meager offering when compared to coeval Moche elite burials (Donnan 1995; Millaire 2002). Despite their unassuming means, the mourners decided to celebrate the death of this woman by performing what is usually described as a

Posts and Pots  3 4 9

“retainer sacrifice” (Besom 2009; Verano 1995) with the ritual killing of another adult female (20–35 years old) whose corpse was also naturally mummified (Burial 9) (figure 13.6). This person was killed as the result of two perimortem blunt force injuries to the occipital region of her cranium, resulting in radiating fractures (Dillon 2013), before her corpse was crammed inside an undersized pit. The force of the blows broke through the ectocranial and endocranial tables of the bone, leaving the right side of the occipital bone and parts of the right parietal, right temporal, and right zygomatic bones separated from the rest of the cranial vault (figure 13.7). Other blunt force injuries were also noted to the frontal bone and nasal area, to the left temporal area, to the right temporal area, and to the maxilla. More than other ritual acts performed during the burial ceremony, the sounds made by the fracturing skull no doubt marked a dramatic visual and auditory performance (see Needham 1967). Carbonized twigs from the ash layer covering the burials provided a

13.6. Burial 9, Room A-102. Photo by Jean-François Millaire.

3 5 0  Millaire

13.7. Cranium of individual in Burial 9 demonstrating catastrophic perimortem injuries. Photo by Jean-François Millaire.

date (Beta-186963) of 1510 ± 50 BP (Cal 430–650 AD), suggesting that the ritual took place sometime during the fifth or sixth century AD. This mortuary ceremony and associated ritual killing no doubt marked a break in how this room was used, but this does not mean that the deceased was forgotten by the families that inhabited the site. In fact, it is likely that this person’s presence among the dwellers was felt and celebrated for years to come through offering renewal, feastings, and other forms of ancestor worship (Millaire 2004, 2015; Nelson 1998; Shimada and Fitzsimmons 2015; Shimada et al. 2004). A Propitiatory Ritual Further excavation in this gallery brought to light traces of an earlier ritual that took place when the room was being built, and its soon-to-be dwellers were possibly concerned with propitiating the forces believed to have control over this place—a type of ritual usually described as a “dedicatory offering” or “foundation sacrifice,” whether the sacrifice involved human or animal blood, food, or inanimate objects (Benson and Cook 2001; Kunen et al. 2002; Mock 1998; Trigger 2003). That the room had held a special status in the eyes of its builders became clear to us when a concentration of ceramic sherds was found below the beaten clay floor,

Posts and Pots  3 5 1

immediately to the south of the first of the three posts that originally supported the roof (figure 13.8). We initially thought this to be part of a fill that served to level the room prior to laying the floor, but we soon discovered that the sherds in fact belonged to two vessels locally known as cancheros that had been broken in situ and intentionally buried. Interestingly, both containers are atypical of the local Virú ceramic tradition. The first (whose body is 17.5cm in diameter) was made with unusually fine clay, featuring red-on-cream burnished stepped motifs enhanced with incisions (figure 13.9). Although clearly Moche-like in terms of its color scheme and decoration, its origin still baffles Moche ceramic experts. It could be a local take on the Moche style, but it hardly fits any local ceramic types identified in the valley so far (Donnan 2011; Ford 1949). The second pot (whose body was 15.8cm in diameter; figure 13.10) is even more remarkable: it fits Virú ceramic canons morphologically but is

13.8. Post 1 with associated sherds. Photo by Jean-François Millaire.

3 5 2   Millaire

13.9. Moche-style canchero with red-oncream burnished stepped motifs (SC8311). Photo by Jean-François Millaire.

13.10. Virú canchero made with kaolin clay (SC8316). Photo by Jean-François Millaire.

13.11. Post 2 with sherds and conical-shaped wooden objects at its base. Photo by Jean-François Millaire.

13.12. Flat-bottomed platter (SC-8320). Photo by Jean-François Millaire.

Posts and Pots  3 5 3

made of kaolin—clay mined from rare deposits in the highland regions of Otuzco and Huamachuco, 70–100km away from Huaca Santa Clara (and 2,500–3,500 meters above sea level). This vessel is also unusual in having extremely thin walls and a polished surface.3 Indeed, although kaolin clay vessels are not unusual in this region, they are systematically Recuay-style vessels that most likely made their way to the coast through long-distance trade networks (see Ford 1949; Larco Hoyle 1962; Lau 2011). This vessel offers convincing evidence that at least some kaolin clay was traded through these channels and that it represented a rare commodity Virú potters sought from their highland neighbors. From this point on, it became imperative to extend the excavation to assess whether this was an isolated feature or part of a larger ritual. To our amazement, another broken container was found next to the second post, together with a number of conical-shaped wooden objects, the function of which is still unclear (figure 13.11). This vessel is a large (27cm in diameter and 6mm thick) flat-bottomed platter with slightly sloping sides and a horizontal rim that is open on one side (figure 13.12). The opening consists of a tong-like surface that extends outward 7cm and whose sidewalls are cut in a step-like fashion. Interestingly, Rafael Larco Hoyle uncovered four similar pieces while conducting fieldwork on the north coast.4 And although the design varies slightly from one platter to the next, they appear to be made of a similar type of clay and have a similar finish. The original function of the vessels is unclear, but the fact that all five pieces share a similar spade-like design suggests that they either served for digging or removing material (but no wear was noted along the opening) or, more likely, for pouring some sort of substance (analogous to Roman pateras). The only other similar dish5 from a known context comes from Barbacoa in the Chicama valley (Larco Hoyle 1941), a site that features a cemetery with Formative Period burials and graves from subsequent periods (Jones 2010). Was the Huaca Santa Clara pouring plate originally manufactured in Chicama, more than 100km to the north? Did it come from a late occupation level, coeval with Huaca Santa Clara, or from a Formative Period burial? Those are difficult questions to answer until more work is carried out at Barbacoa. As the excavation continued, we were mildly surprised to find that yet another vessel had been broken next to the third post before the beaten clay floor was installed (figure 13.13, figure 13.14). Unlike the other ceramics, however, this is not a foreign object but rather a classic Virú canchero (the body is 17.9cm in diameter) that features typical

3 5 4   Millaire

13.13. Post 3 with associated sherds and rocks. Photo by Jean-François Millaire.

negative decoration (Millaire 2009a). Cancheros of this type are common at Huaca Santa Clara and indeed at other Virú sites, including the Gallinazo Group. What is surprising is not the container per se but the fact that the rock used to crush the vessel was left in place. Sitting on top of the broken dish, its underside was still covered with baked clay dust. Interestingly, fragments were missing for all vessels mentioned above except for the Virú canchero, suggesting that the containers were either not broken on the spot or, perhaps, that pieces were collected by attendants, either to keep or to be used as pars pro toto (or surrogates) of complete vessels in subsequent rituals (Hecker and Hecker 1992).

Discussion Why crush these pots below the floor of a gallery? The only line of evidence we have is that these were no ordinary containers, including an exceptional Moche-style piece, a unique canchero made with imported kaolin clay, and a possible libation platter. What kind of message were these offerings meant to convey to the sacred order or to those present? Something related to alliances with neighboring and distant polities with which they had established trade relations? It seems a likely hypothesis—but a very hard one to test.

Posts and Pots  3 5 5

On a different level, in his introductory paper to a World Archaeology issue on this form of ritual, Osborne (2004) asks, “Why does votive deposition matter?” It matters, he argues, because the exchange of material objects for supernatural returns was both socially and economically significant for the societies that performed these rituals. Or, as Cook (2001:139) puts it, one should in fact think of such offerings as a form of “ritual tribute” to the divine. In some cases, foundation sacrifices were aimed at establishing a particular model of reciprocity with supernatural powers, while in other cases they were clearly meant to consecrate a building or turn “secular ground into a sacred ground” (Marcus and Flannery 1994), a process Blom and Janusek (2004) and Swenson (2012) appropriately describe as “place-making.” Dedicatory sacrifices of this type are also not unusual in other parts of the world, where they range from wrapped stones to shells, broken effigy jars, and human beings (Kunen et al. 2002; Mock 1998; Osborne 2004; Trigger 2003). For example, in Southeast Asia, traces of foundation sacrifices are often uncovered near the architectural base of temples or palaces, where they are interpreted to have propitiated the gods

13.14. Virú negative canchero (SC-8325). Photo by Jean-François Millaire.

3 5 6   Millaire

or the ancestors. On Nias Island in Indonesia, traditional village rulers’ houses were exceptional not only because of their disproportionate size and complexity but also because trophy heads were usually buried near the foundations (Feldman 1977; Scarduelli 1990). Dedicatory offerings are also conspicuous throughout Mesoamerica, where they represented one of the most prevalent ritual activities (Kunen et al. 2002; Mock 1998; Morlion 1999). At Piedras Negras, for example, bowls containing precious or exotic objects, including eccentric flints, obsidian blades, jadeite pieces, and shells, were placed in the ground before the construction of Structure O-13 (Coe 1959). Similarly, at San José Mogote, Marcus and Flannery (1994) note that exceptional artifacts were found inside offering boxes beneath the floor of a temple where they had been buried following rituals of sanctification. Propitiatory sacrifices are also common in the Andean region of South America (Benson and Cook 2001; Kunen et al. 2002; Mock 1998; Osborne 2004; Trigger 2003). At the Late Preceramic site of El Paraíso, for example, Engel uncovered a small stone structure close to Unit I that Quilter (1985) has argued functioned as a dedicatory offering. It contained a stone covered with red pigment, wrapped in a textile, together with a gourd bowl, food, and a miniature fiber bag—a small version of those filled with rocks and used for room fill. Similar early foundation sacrifices are reported from Kotosh, where an offering of three headless bodies was found underneath the Late Initial period temple (Burger 1992a:118), and from La Galgada, where a dedicatory offering of shell disks wrapped in textile was uncovered at the foot of the main staircase (Burger 1992a:120). Proulx (2001) notes that Nasca iconography portrays human trophy heads being entombed underneath pyramid-shaped mounds, and indeed numerous trophy head caches have been found in this region. At Conchopata, archaeologists uncovered the remains of a temple foundation ritual in which five young females and more than 20 oversized anthropomorphic jars were “ritually killed” (Cook 1987; Isbell 1987; Isbell and Cook 2002). At the base of the Akapana, in Tiwanaku, archaeologists found extensive evidence of dedication sacrifices involving humans as well as large numbers of polychrome vessels (Blom and Janusek 2004). At Huacas de Moche, many human effigies were uncovered in Plaza 3A and Plaza 3C with vessels broken in situ as offerings, perhaps as substitutes for human lives (Bourget 2001b; Verano 2008a). Compelling cases of dedicatory sacrifices were discovered at the Sicán archaeological

Posts and Pots  3 5 7

complex in the Pomac forest of the northern La Leche drainage. Ritualized burial and paired offerings of objects and people atop the multiple huacas at Sicán, along with a unique seated burial of a young child to the front of Platform II at Cerro Cerrillos, seem to carry on this tradition during the opening of the Late Intermediate Period (Shimada 1995; Klaus et al. 2010). And at Chan Chan and Pacatnamú, dedicatory offerings of young women were commonly placed under doorways and ramps of royal palaces (Verano 2008b). But to what extent and in what ways are the human sacrifices at Conchopata, Tiwanaku, and Sicán different from the sacrifice of objects such as the exceptional ceramic vessels at Huaca Santa Clara? I would venture to say that these rituals fall into two distinct categories and that the chasm is deep enough to justify singling out human sacrifices as a research theme. This is rather strange considering how the well-documented Inka classified sacrificial offerings (Rowe 1946). Although human beings were the most valuable, they were the costliest in a long list of worthy sacrifice items, including animals, food and drink, and inanimate objects, including textiles. Although Inka religious practices were probably as far from north coast realities as were, say, La Tolita Culture beliefs, this should minimally make us wary of the dangers of myopically overemphasizing human remains when discussing a deeper and more complex concept of ritual killing and offerings. In my mind, the difference between the dedicatory offering of oversized jars at Conchopata, versus the ritually killed column holders and other inferred dedicatory burials at Huacas Lercanlech, Loro, and Las Ventanas, is simply one of scale and a reflection of a conceptual continuum (as is the difference between shattering special vessels at Huaca Santa Clara and shattering a retainer’s skull later on in the room’s history). In the absence of universals to assess the relative value of these acts of relinquishment, I believe we should place the performers at the center of the rituals rather than focus on what is being surrendered. To build on Becker’s (2001) words, foundation sacrifices carried out in major public buildings versus those performed in more modest places are likely to differ only in the lavishness with which they are enacted. In this context, it seems that sacrifices are not only and always proportionate with the needs of those who wished to consecrate a building, position themselves in the socio-political exchequer, or maintain a relationship with the sacred order; in fact, they usually are a reflection of people’s means at a specific junction in time and space.

3 5 8   Millaire

Conclusion The ritual killing of objects in the Andean region was probably much more prevalent than we have traditionally thought. Crafted items of all sorts were doubtlessly commonly destroyed as part of acts of worship to establish, maintain, or restore a necessary relationship between humans and the sacred order (Faherty 2011). This leads to important questions: How many killed objects might we have seen before but perceived only as broken pots? Should we pursue an “archaeology of object sacrifice” to discriminate between ritual practice and trash disposal? In any event, we should not over-theorize this issue. Sound excavation techniques coupled with an awareness of the rich, complex, and changing interactions between individuals and groups will be key to future answers. Acknowledgments I would first like to thank Haagen Klaus and J. Marla Toyne for inviting me to what has been a very productive symposium and subsequent writing project. This research was made possible thanks to the financial support of the Fonds de Recherche du Québec—Société et Culture and Canada’s Social Sciences and Humanities Research Council. My warmest gratitude goes to Estuardo LaTorre Calvera, Jeisen Navarro Vega, and David Chicoine for their valuable contributions to the Huaca Santa Clara research project. I would finally like to thank Magali Morlion for her patience and unwavering critical thinking and the editors and reviewers for stimulating comments on earlier versions of this chapter.

Notes

1. The term Virú refers to the state-organized society that developed in the Virú valley during the Early Intermediate Period. First described by Rafael Larco Hoyle (1945), this society was later referred to using the problematic term “Gallinazo” (see Millaire 2009a). 2. Maize, beans, squash, guayabas, peanuts, pacae, avocados, lúcuma, cotton, sweet potatoes, and peppers were identified, among other crops. 3. A similar vessel can be found at the Museo Larco (ML006408) but in this case it is a red-ware vessel covered with a white slip. 4. See catalogue numbers ML016940, ML016941, ML016942, ML017118 (museolarco.org). 5. Catalogue number ML017118 (museolarco.org).

Practicing and Performing Sacrifice  36 1

chapter fourteen

Practicing and Performing Sacrifice Tiffiny a. Tung Department of Anthropology, Vanderbilt University

The studies of sacrifice on the north coast of Peru presented in this book expand our knowledge of the region, particularly as to how the bodies of sacrificed humans and animals were treated. This has provided a clear picture of those practices and thus has great potential to clarify how sacrificial acts were integrated into other areas of domestic, political, and social life. Through these studies, we gain a sense of how rituals of sacrifice could create, reinforce, or challenge particular social and gender norms, roles, and hierarchies, among other facets of society. Indeed, that is one of the strengths of this contextualized approach, wherein the detailed analyses of skeletal remains are used to reconstruct ritual processes and interpret how those practices are connected to other facets of an individual’s life and society at large. The analytical strength of archaeology and bioarchaeology is the focus on the material and the spatial, which gives us a closer and more clarifying view of people’s practices and minimizes the obsession with getting in the minds of the skeletons we study. This volume does much to show those strengths through detailed studies of sacrificial victims (humans but also animals and objects), the places where the events occurred, and the associated practices that together constituted a sacrifice. That is, although many of the authors

3 6 2   Tung

did not explicitly discuss it this way, their empirical studies of the physical transformations made during sacrifices bring us closer to realizing an “archaeology of sacrifice/ritual/religion”1 that focuses on practices, not a hand-wringing about the inaccessibility of beliefs among a people that are long dead. I see this as evidence of maturation in our discipline and an analytical tack that is sometimes in line with a practice theory approach, in which we describe the actions of people and evaluate how those actions may have been influenced by personal experiences (e.g., evidence of previous violent encounters), historical contingency (e.g., long-term conflicts with neighboring regions), and larger social forces (e.g., norms or unspoken rules about what young men or old women should or should not do). In this context of trying to use archaeology to get at practice, I agree with Campbell (2012:308) when he says it is a misunderstanding that archaeology is not well suited for studying ritual/ religion due to the “Protestant-inflicted notion of Modern intellectualist scholars of religion that religion is about belief rather than practice.” This volume on sacrifice has aided in clarifying three key issues, certainly within the setting of pre-Hispanic north coast Peru and, perhaps, more broadly in other spatial and temporal contexts: (1) definitions of sacrifice and the near-consensus that it should be conceived as part of religious practice, or what many of the authors view as a means for communing with deities and ancestors; (2) the theoretical framing of sacrifice and the common division between the sacred and profane (mundane) to identify and distinguish sacrificial acts from other activities (despite some longstanding criticisms in anthropology that challenge this Western dualistic approach); and (3) theoretical approaches that evaluate the structuring effects of sacrifice. That is, how does sacrifice affect the social and political realms, particularly as it relates to social and gender roles, hierarchies, and the organization of society more generally? And simultaneously, how does a society itself structure the practices of sacrifice? This volume aids in identifying concrete examples that clarify how sacrifice structured, and was structured by, society. For example, we have an exploration of how Moche gender roles affected who was selected for sacrifice and who participated in the ritual acts (Verano and Phillips, chapter 9); a discussion about how conceptions of personhood and group identity shape how a group might select and ultimately process the sacrificial victims (Klaus and Shimada, chapter 5); and a stimulating argument about how the ritual destruction of ceramic objects shaped the ongoing practice

Practicing and Performing Sacrifice  36 3

of ritual sacrifices (in all its forms), which were meant to foment and continue relations with the supernatural (Millaire, chapter 13).

What Is Sacrifice? In this book, humans, animals, and objects are all given analytical treatment in the discussion of sacrifice in northern coastal Peru. This strikes me as an intellectually stimulating approach, and I will not delve into the long-standing debates about the animacy and agency of animals and inanimate objects, something others have discussed elsewhere (Ingold 2007; Latour 2005; Tung 2014; Williams 2004). Further, because the authors who discuss the sacrifice of animals (Gaither et al., chapter 6; Szpak et al., chapter 12) and objects (Millaire, chapter 13) explain how they are part of the larger Andean practice of offering meant to nurture relationships between humans and the divine, I prefer to focus on the definition(s) of sacrifice (or lack thereof ) within the volume to highlight the parallels and distinctions in how sacrifice was identified, described, and analyzed. In chapter 1, the editors note that this book speaks to the challenges surrounding archaeological definitions of sacrifice, and they present a nice summary of the various ways that previous scholars have defined sacrifice. The editors do not set forth a definition that the contributors had to follow. Instead, each author approaches the topic independently, some clearly articulating how sacrifice should be defined, understood, and studied; others marched on with the assumption that there was agreement as to what constituted sacrifice and what role it played in ancient pre-Hispanic societies. Should sacrifice be placed squarely within the realm of religion (Schwartz 2012) because, as some argue, it is about communing with the supernatural and “affecting the suprahuman realm of immaterial entities” (Klaus and Toyne, chapter 1 citing Tatlock [2006])? Or, as Klaus and Toyne suggest, might that be too limiting because it excludes ritualized executions on the more secular end of the spectrum, as in the cases of ritually executed war prisoners (Verano 1986a). Does that thus require a distinction—ritual sacrifice versus secular sacrifice—or are those categories more reflective of our post-Enlightenment “intellectualist reification of ‘religion’ as a unique and independent sphere of practice and (especially) belief” (Campbell 2012:306), combined with an

3 6 4   Tung

intellectual embodiment of the US Constitution’s First Amendment (at least for American scholars)? As Campbell (2012) argues, sacrifice should not be distinguished and embedded in “the sacred” or in “religion” because in many ancient and modern societies religion is inextricably linked to politics and to the social. Building on Asad (1993), he notes that the modern West has come to view “religion” as a perspective and, thus, as an analytical category into which the study of sacrifice is placed. For Campbell, using the etymology of the word sacrifice (to make sacred/holy) as a starting point essentially privileges the Western conception, and he posits that this etymology is a “historically positioned construction that may carry unwanted baggage into translocal comparison” (2012:306). He continues by noting that the division of the sacred and profane, so favored in Western intellectual discourse, “has even motivate[d] some theories of sacrifice,” such as Hubert and Mauss (1964 [1898]), that are commonly used to explain a wide variety of sacrificial practices. While I am sympathetic to that view and recognize that our scholarly insights can grow from appreciating such a non-dualistic perspective, we cannot assume that pre-Hispanic populations in the Andes were one way or the other. Was religion or ritual inextricably intertwined with all other aspects of community and individual life, such that we cannot identify and describe distinct ritual acts? Or was “religion” somewhat demarcated from other social spheres such that practices could be experienced and perceived as ritual or not? As the data accumulate, and as this book does much to bring a vast array of empirical datasets together, we can describe and evaluate how practices of sacrifice are demarcated from or intertwined with other areas of activity while simultaneously exploring how they affect and are affected by other aspects of domestic, political, and social life.

Defining Sacrifice Four authors explicitly define or discuss definitions of sacrifice in their chapters. Chicoine (chapter 4) highlights the “preprogrammed repetitive action” of sacrifice (citing Benson and Cook [2001]) and the requirement that something important be given up “for a reason perceived to be of greater importance in some manner than what is to be sacrificed” (Chicoine, citing Green [2001:19]). This is relevant in his discussion of strangulations in the southern sphere of the Moche realm, in which humans

Practicing and Performing Sacrifice  36 5

are ritually strangled perhaps to control the “supernatural, natural, or even sociohistorical forces” (Chicoine, chapter 4), in a manner distinct from the ostentatious sacrifices of bloodletting and dismemberment that may have metaphorical ties to feline predators. Indeed, he argues that ritual strangulation may represent an inversion of the “sacrificial values of spectacle,” revealing a more “restrained, controlled, and quiet form of killing” that might have metaphorical connections to the constricting killing methods of some snakes. Nonetheless, it is still unclear how preprogrammed and repetitive these strangulation sacrifices were, something acknowledged by Chicoine and to be clarified with additional studies at Pañamarca and other southern Moche sites. Tomasto et al. (chapter 11) tackle the definition head-on, accepting the key point that sacrifice involves communion “with the realm of the sacred,” but offerings of sacred objects (e.g., isolated bones; ceramic pots [Millaire, chapter 13) are not sufficient; something or someone must be destroyed or killed. And the majority of the chapters hint at or explicitly acknowledge in some way that destruction or violence against the entity being offered was a requisite for constituting sacrifice. Indeed, “Sacrificial slaughter involves violence” (Aldhouse-Green 2001), and this dramatic transformation from a whole to a fractured pot, or from a living entity to a lifeless sacrifice, often entails mutilation, fierce beatings, and other physical acts far beyond that required to destroy the ceramic vessel or kill the person (Klaus and Shimada, chapter 5; but see Chicoine, chapter 4). It is a form of “overkill” that can be detected by researchers as they examine cut marks and perimortem fractures on intact skeletons or pieces of bone or impact scars and breakage patterns on smashed pottery (Millaire, chapter 13). Thus, in the pre-Hispanic north coast of Peru, some form of corporeal modification or outright destruction was a very common component of sacrifice rituals. Szpak et al. (chapter 12) remind us that animal sacrifice and the “killing of an animal under ‘ordinary’ circumstances” can be highly ritually charged, and for some cultural groups all animal kills are considered sacrifices. Yet they advance their study by clearly articulating that they identify “animal sacrifices” as those in which animals are slaughtered in “highly ritualized contexts” in which the ritual is “not incidental but is foregrounded.” Drawing from Ingold (1987), among others, they further note that “sacrifice (rather than simply offering) implies the giving of something that is owned by the donor,” as in the herder-animal relationship as opposed to the hunter-hunted relationship (Ingold 1994).

3 6 6   Tung

However, one could argue that the immense time, skill, and energy required to hunt down an animal that is subsequently sacrificed may warrant consideration not just as an offering but as a sacrifice. Szpak et al. (chapter 12) cogently argue that the human-animal relationship significantly differs for those engaged in animal husbandry versus those who hunt, including for those who do both. Domesticated animals may be more overtly recognized as stand-ins for humans, and people would have certainly known some of those domesticated animals quite well. Gaither et al. (chapter 6) discuss the practice of sacrifice in terms of its intended functions or desired outcomes. Drawing on Bourdillon’s (1980) ideas about sacrifice, they note that “calculated sacrifices” may be a form of gift-giving to create lines of communication and reciprocity between the living and the supernatural or the ancestors. This framework is often cited in contexts of environmental stress, other natural disasters, or requests from “nature.” For example, periods of extended drought or torrential floods, earthquakes, or poor harvests may require human and animal sacrifices to obtain “goodwill from the gods” (Gaither et. al., chapter 6). But more to the point, this acknowledges that sacrifice is about communing with the supernatural, often to receive benefits, a perspective that Aldhouse-Green (2001) expands upon, arguing that sacrifice is not an end in itself but a means to fulfill a request. While Aldhouse-Green’s perspective may be accurate in some cases, I think that it fails to recognize the practice of sacrifice itself and how it transforms not only the one who is sacrificed but also those who produce it and observe it. Those transformations may thus be related to what Bourdillon (1980, as cited by Gaither et al.) notes as “calculated sacrifices” (discussed above), “prestige sacrifices” that aggrandize the status of those responsible for the sacrifice, and sacrifices for “sociopolitical cohesion,” though none of those are mutually exclusive.

Theoretical Framing of Sacrifice “Sacrifice is a religious act that can only be carried out in a religious atmosphere and by means of essentially religious agents. But, in general, before the ceremony neither sacrifier nor sacrificer, nor place, instruments, or victim, possess this characteristic to a suitable degree. The first phase of the sacrifice is intended to impart it to them. They are profane; their condition must be changed. To do this, rites are necessary to introduce them into the sacred world and involve them in it” (Hubert and Mauss 1964 [1898]:19).

Practicing and Performing Sacrifice  3 67

Most chapters in this volume either implicitly or explicitly use Hubert and Mauss’s theoretical framework in which distinct spheres of religious and secular activities are recognized and sacrifice is perceived to be within the realm of religion. Many of the studies evoke explanations of communing with the supernatural, often in an effort to bring about some desired outcome as it relates to the natural world. Whether a group was sacrificing a human, an animal, or an object, all of these could be conceptualized as “ritual tribute” to the deities and ancestors (Cook 2001:139). One of the chapters that challenged that approach, or at least viewed the ritual killings as distinct from being a religious act, is the one in which “captive executions” were distinguished from blood sacrifices involving “the offering of human lives for religious purposes” (Verano and Phillips, chapter 9). The broader anthropological debates about notions of religion and secularity are brought forth in these archaeological studies of ritual killings. The detailed descriptions of how the sacrificial acts were conducted, the life histories of those who were sacrificed, and the larger environmental and sociopolitical circumstances that ritual killings occurred within provide the empirical framework to theorize about sacrifice and its relationship to religious organization (e.g., did only a specific class of ritual specialists carry out the sacrifices, resulting in a highly standardized method of ritual execution?), community organization (were there markedly distinct social classes that had distinct roles?), social structure (were gender roles clearly demarcated and differentiated?), and political structure (was military prowess a means for political authority?), among other aspects. In Moche times, gender roles may have been a powerful structuring force in determining who could be captured, ritually executed, and dumped in the plazas of huacas; in this case, male identity was an integral factor in their selection as victims (Verano and Phillips, chapter 9). Further, their apparent social role as warriors further marked them as ideal for this kind of ritual execution; it may have been that not just any man (for example, elderly men are not a part of this executed group), but the identity of “male warrior” is what marked one as appropriate for being captured and ritually executed. If we take Hubert and Mauss’s (1964 [1898]) and Aldhouse-Green’s (2001) definitions and expectations of what constitutes sacrifice, then the ritual killing of war captives may not be sacrifices per se; those ritual killings may be an end in themselves and represent the culmination of a war campaign, revenge killing, or some other event. That is not to say that the ritual killing had no broader implications for society; to the

3 6 8   Tung

contrary, the ritual killing and the attendant ceremonies may all serve to maintain social roles and hierarchies or to challenge existing ones. More specifically, as with those ritually executed at Huaca de la Luna, Pacatnamú, and Punto Lobos (Verano and Phillips, chapter 9), those acts may be more directly tied to military and territorial expansion. Communion with the divine may or may not have been sought.

Structuring Effects of Sacrifice and Other Violence The detailed descriptions of the ritual killings—from cut mark patterning, to skeletal element representation, to descriptions of ante- and perimortem trauma, to the dietary reconstructions of the animal and people prior to their deaths—provide us with a rich understanding of how the actual sacrifices were conducted. Beyond that, several of the chapters also examined how those sacrificial acts shaped and were shaped by other social and political forces. In comparing these sacrifices on the north coast of Peru to other areas of the Andes—the Wari in particular—it appears that there were some shared processes and effects in the Moche and Wari societies. For example, Wari society was quite similar to Moche society as it related to the ritual killing of war captives, but it diverges in that the Wari subsequently transformed their executed dead into trophy heads (though there is evidence that the Moche did this on occasion) (Verano 1995). These particular practices structured Wari social and political organization in profound ways. For example, the capture and sacrifice of war captives likely aided in establishing and enhancing the Wari military apparatus and notions of military supremacy while also providing a path to power for certain groups of men (those who engaged in the military excursions and captured victims) (Tung 2012). The Wari trophy heads were eventually deposited in unique ritual structures (D-shaped buildings), a practice that reveals the inextricability of Wari ritual and military activities and how they mutually constituted each other. The act of sacrificing war captives, ritually transforming them from living humans into trophy objects, and artistically representing those acts on polychrome ceramics and textiles, also required the skill and knowledge of various specialists. In short, these ritual killings and the attendant ceremonies served to create and maintain new social roles: military specialists tasked with obtaining captives (Tung 2012), ritual specialists who created trophy heads (Tung 2008), and master artisans who produced the wares with the pictographs of bound prisoners,

Practicing and Performing Sacrifice  36 9

warriors with weaponry, and human trophy heads (Ochatoma and Cabrera 2002) were all involved in the process. With each new prisoner captured, each new trophy head made, and each new ceramic vessel produced, those social roles were reinforced, showing how many different aspects of Wari life were deeply interwoven and constantly reinforcing each other. It appears that the Moche also had these similarly structuring connections between military and ritual life, though perhaps certain ritual killings should not be perceived purely as religious in nature. Another example of how ritual practices might reflect and create particular sociopolitical formations can be seen in the discussion of ritual strangulation. Chicoine (chapter 4) argues that the “nonspectacle” aspect of ritual strangulations, practiced in the southern Moche sphere, may have been a means for performing and maintaining the sociopolitical and religious differences between the northern and southern regions of the Moche world. The selection of particular individuals as “retainer sacrifices” may reveal insights into gender roles and the role of kinship in ancient societies. For example, the greater frequency of female retainer burials during and after Moche times (Bentley and Klaus, chapter 10) is an important data point that may allow us to better comprehend how women were perceived and what their social expectations were. In that same chapter, the authors also conducted biodistance analyses to explore how kinship may have structured who was selected for retainer sacrifices. Although some of this work is still in its early stages, it is an exciting line of inquiry because it recognizes the strong force that biological kinship might have in shaping social identity and thus stipulating who might be appropriate for sacrifice. As those norms are incorporated into society and then shape who gets selected for retainer sacrifice, it has the cyclical effect of reinforcing the importance (or lack thereof ) of biological kinship ties, particularly as it relates to performing “proper” burial rites and communing with the ancestors in that process.

Conclusion: Ritual Violence and Violence While sacrifice and ritual violence may have broader implications for the social and political, there are forms of violence that are not necessarily linked to the ritual. For example, in Wari settings (Tung 2012), Tiwanaku-affiliated sites (Torres-Rouff 2011), Colonial Arequipa (Chambers 1999), modern Bolivia (Van Vleet 2002), and other contexts (Counts et al. 1999), the broken noses and other trauma on females

37 0  Tung

may have more to do with social and gender norms that see physical abuse of women as an acceptable practice. In these particular contexts, the intra-household violence—whether perpetrated by male partners or other “superiors”—may be disconnected from ritual violence. But these forms of physical violence may be partially created and sustained by a larger milieu of violence in the community—ritual killings, for example. Recognizing the various cultural practices, from the mundane to the sacred, that serve to perpetuate violence in all its forms is an analytical task that has been approached by a variety of researchers. Thus, although Klaus and Toyne (chapter 1) note (citing Klaus [2014b]) that on the Peruvian north coast there is “a relative absence of widespread evidence associated with interpersonal or sociopolitically sanctioned violence,” I note that in other places and times there are correlations between ritual violence, the material propagation of that violence (e.g., iconography or architectural spaces that condone and celebrate the violence), militaristic violence, and violence in other contexts (e.g., intra-community violence in which men, women, and/or children are physically beaten [Gaither and Murphy 2012]). How is it that in some societies violence is both so mundane and so highly ritualized, as in Wari society in which we see female victims of sublethal domestic or community abuse, and men and children with fatal trauma, mutilation, and eventual transformation into trophy heads (Tung 2012)? Others might even argue that this is the case in contemporary US society, where there are highly trained soldiers engaged in routinized violence that often bleeds into the mundane sphere of domestic life (Lutz 2007; Nordstrom 1998). In contrast, there are other settings—the Moche world, perhaps—where sacrifice and ritual violence are highly circumscribed, both in space and victim profile. Granted, additional archaeological samples are still needed from Moche contexts to evaluate how pervasive violence really was, but books such as this one raise important questions beyond that of sacrifice and implore us to examine violence in all of its manifestations. Note 1. I have grouped together these terms because all but one of the authors in this volume discussed sacrifice within the context of ritual and religion (or at least communion with the deities or other supernatural forces).

Mesoamerican Perspectives  37 1

chapter fifteen

Mesoamerican Perspectives on the (Bio)archaeology of Andean Ritual Violence Vera Tiesler Facultad de Ciencias Antropológicas, Universidad Autónoma de Yucatán, Mexico

Multilayered archaeological social phenomena, such as ritualized violence and specifically human sacrifice, can be studied and understood culturally only after combining scientific approximations and confronting different lines of hard and discursive evidence. Scholarship across the globe has approached the near universal subject of human ritualized violence from a myriad of angles, grounded in the zeitgeist of each epoch, scholarly traditions, and, naturally, the information at hand. There are academic specificities for each cultural area, and the anthropology of Peruvian ritual violence is no exception (Benson and Cook 2001), just as this volume illuminates the study of its north coast cultures. Here, I will examine the scope of regional scholarship and then reflect upon it briefly from a “view from afar” within the anthropology of ritual violence in Mesoamerica. In doing so, I wish to shed light on some of the agendas, methods, and challenges faced by this line of research in the Andes and, at the same time, encourage potential future avenues of exploration.

Studying Ritual Violence on the North Coast of Peru The endeavor of studying ancient ritual killings is especially challenging in those societies of the past that lack an eloquent written record,

37 2   Tiesler

thus leaving the material registry as the primary source of information for cultural reconstruction and interpretation. This is the case in those pre-Columbian societies that once occupied the Peruvian north coast. Here, ritualized violence has been made known only tangentially in the broader ethnohistoric or ethnographic record. Pictorial information and archaeological reconstructions have been, by far, more central to understanding this aspect of regional history. Bioarchaeology and human taphonomy—and more recently archaeothanatology (after Duday 2009)—have become prominent venues for exploring ritual violence in the area. These approximations blend mortuary archaeology with physical anthropology and forensic orientations, a combination ideally suited to untangle the complex mortuary pathways that stem from sequenced acts of body processing has been demonstrated in other areas of the Western Hemisphere (Benson 2001; Boone 1984; Chávez 2010; López Luján and Olivier 2010; Tiesler and Cucina 2007; Turner and Turner 1999; White 1992). On the north coast of Peru, these inquiries made upon human assemblages have received a strong boost since the late twentieth century through a series of germinal archaeological discoveries at the ancient sites of Pacatnamú, Sipán, and Huaca de la Luna by Steve Bourget (2001b), John Verano (1986a, 1986b, 1995, 1997, 2001a, 2001b, 2007), and colleagues. Many other researchers have followed in their footsteps in recent years, most of them interested in the bioarchaeology of human remains (Anderson-Hamilton 2005; Gaither et al. 2008; Klaus 2008a; Toyne 2008). It is the principal goal of this timely volume to unite and update all the data that have accrued over some three decades at Peruvian north coast settings and understand these jointly within their dynamic cultural and social undercurrents. Other goals as expressed by Klaus and Toyne (chapter 1) revolve around a context-embedded research (suited to tie together the strings of different data sets, times, and cultural scenarios); the incorporation of new techniques and procedures (such as quantitative approaches in taphonomy, cut mark microscopy, life history approaches to paleopathology, isotope research, and so forth); and relevant social concepts that encompass the phenomenon of sacrifice from an expressly holistic view. For this, each contributor brings in his or her own working hypotheses and in-field observations, ideas, and notions, as the flow of chapters nicely demonstrates. Regarding the data and reconstructions these authors develop for the north coast of Peru, the sheer complexity and diversity of the documented

Mesoamerican Perspectives  37 3

pre-, peri-, and postmortem body treatments and the myriad forms of death, depositions, redepositions and reuses are truly puzzling. The authors identify exsanguination, strangulation, slaying, butchering, or possible poisoning in the documented human assemblages. Some other victims appear to have been buried alive. Related body-processing appears to have included heart ablation, defleshing, dismemberment, forced joint dislocation, recycling of body parts, trophy acquisition, and perhaps flaying. Still other forms of protracted mortuary behaviors, such as prolonged surface exposure and mummy transport, seem to completely blur conventional taxonomic categories of primary versus secondary burials and the duration of burial rituals. As the editors and most of the contributors underscore, this diversity in human body-processing, much of which is presumed to be ritual, demands tightly imbricated lines of argumentation and extensive use of refined forensic patterning. The detailed maps, skeletal drawings, and identification of bodies and body-part manipulation show a degree of refinement that speaks not only of the care taken during excavation but also of the excellent preservation of most documented human assemblages. They render credibility to the overall scenarios that culminate in each contribution and invite broader questions beyond the sites of killing. One might wonder what happened to the many missing body segments and isolated body parts at Huaca de la Luna, for instance (Hamilton, chapter 2; Backo, chapter 3). Were there specific meanings attached to differing body parts? Did any of those missing pieces undergo recycling as raw material for crafting bone artifacts or trophy heads (i.e., Verano et al. 1999)? Or were they discarded in the site’s middens? What were the normal, reverential burials like at each settlement, and how did they differ from the patterns documented in this volume? Indeed, the protracted nature and the diversity of treatments presented in this book are met only by a broad approach that encompasses reconstructions of complete mortuary pathways, an aspect that sets the scale for most chapters in this volume. This was the approach originally pursued by John Verano at Pacatnamú, Sipán, and Plazas 3A and 3C at Huaca de la Luna, to which Hamilton (chapter 2), Backo (chapter 3), and Bentley and Klaus (chapter 10) add information and hindsight. The human record of sacrifice or suspected ritual violence (the so-called grey area) is put into perspective by confronting “deviant” versus modal patterns. These assessments go laudably

374  Tiesler

beyond simple counts of anthropogenic marks on bones and skeletons. Instead, those who are presumed to be victims of retainer, offering, or foundation immolation are varyingly examined according to their health, provenience, gender, age, and biogeochemistry. Jointly, what could be disparate lines of data combine to provide points of departure for social reconstruction. Local folk are distinguished from foreigners, and lifestyles are assessed from skeletal markers to define the marginalized, the mainstream, and the mighty. Still more specific social identities are ascertained, such as those consistent with warrior status (or not). The latter, for example, constitutes the main focus of Verano and Phillips’s (chapter 9) convincing confrontation of healed or healing fractures among sacrificial victims from Moche and Chimú sites. Beyond methodology and behavioral reconstruction, there remains, of course, the challenge of inferring the specific meanings and broader sociocultural undercurrents that once motivated the ritual killing suspected from the archaeological record. I was glad to see that the discussions in the chapters go beyond the categorical notion of “there was”/“there was not”; and instead make more nuanced arguments of possible forms of death surrounding behavior, cultural meanings, and undercurrents in each case. And here lies probably the main challenge for this volume and of regional (bio)archaeological research in general: translating the archaeological and mortuary records into meaningful sequenced behavioral patterns and ritual programs, an endeavor that the contributors have attempted quite successfully by integrating broader archaeology, iconographic interpretation, and ethnohistoric research. Combined with skeletal data sets, many chapters distinguish culturally aligned patterns and recognize potential sacrificial forms and related treatments, such as child immolations, retainers, “like-with-likes,” and re-interment of body parts. The inferred taxonomies and the reconstructed mortuary pathways are directly relevant to the role of human sacrifice (which might be described as the ritual killing and consecration of a person) within a complex social and religious setting. In all likelihood, sacrifice followed prescribed liturgies, and arguably just one (albeit culminating) element of a given religious program. Other steps in the unified native liturgy should have included prayer, offerings, pre-occasion rites, invocation, donation, or termination. The enactment of these steps most probably shared a conservative quality and likely involved additional or analogous material elements, as Steve Bourget (2006) has previously argued

Mesoamerican Perspectives  37 5

regarding Huaca de la Luna ceramic vessels, or what now has been proposed for foundation and dedicatory rites in general. Unfortunately, the dearth of discursive information for coastal Andean ideological expression(s) or ethnohistorical testimonies limits such interpretive attempts. Moreover, the bewildering diversity of body-processing per se, some clearly lacking any cathartic element, comes to challenge any ad hoc understanding of what could be viewed as religiously motivated human sacrifice in its strict sense (Beattie 1980; Bell 1997; Hubert and Mauss 1964 [1898]; Rappaport 2004) and what was not. Note that scholarship relates human sacrifice and, to a lesser extent, post-sacrificial body treatments to an extreme form of mortuary expression: that of institutional and highly redundant, conservative ritual liturgies of consecration, usually with a violent, kratophonous (destructive) element involved and enforced by authorities in public or sacred spaces. The variation in body treatments that resonates across this volume calls into question simplistic, one-dimensional interpretations. Instead, the authors make clear that multiple meanings played out among the sacred rites along the north coast of Peru. And, as some of the authors suggest, some of the archaeologically reconstructed acts of interpersonal aggression should express not institutionalized sacrifice and ritual violence but only different forms of “nonritual” physical assaults against others.

(Bio)archaeological Perspectives on Ritual Violence in Mesoamerica Many issues voiced in this volume for the Andean sphere also find resonance among Mesoamerican scholars. I draw attention, for instance, to the challenges Mesoamericanists face in assigning potential sacrificial status to human assemblages from cached remains, isolated bone scatters, or primary inhumations. The first category identifies mostly intermingled and incomplete human assemblages in clearly offertory arrangements that have habitually been labeled as “caches,” thus explicitly or implicitly identifying clusters of human remains as ceremonial artifacts—a notion essentially distinct from the ancestral statuses implied by funerary disposals. Some problems associated with this dualistic nomenclature, which simplified the multiple origins of human assemblages, have been acknowledged by William Coe (1959, 1965), who established this system for the Lowland Maya. More recently some authors, such as Marshall Becker, have come to view human assemblages not through reverential or sacrificial dichotomies but as part of a broad

37 6   Tiesler

continuum of earth offerings, with some burials appearing in the form of a cache and others as caches assembled as with a burial (Becker 1992, 2001; also see McAnany et al. [1999]). This spectrum also includes miscellaneous human remains or “problematic deposits,” a category made up largely of random bone scatters in the fills and construction middens of Mesoamerican sites. This “problematic” label admits the impossibility of assigning defined cultural meaning to isolated human components beyond suggesting a series of potential formation processes that could relate to ritual or domestic refuse, recycling of skeletal elements in the manufacture of goods, or protracted funerary practices. Were these severely commingled assemblages to stand for post-sacrificial discard, where they would, beyond doubt, speak for the importance, complexity, and length of ritually determined forms of body-processing? In recent years, these concentrations have been approached variously in terms of their depositional histories, presence in symbolic spaces, or underlying human agency associated with ritual qualities and the vital essences still believed to reside in specific body parts or bones (Chávez 2010; Mock 1994; Kunen et al. 2002; Talavera et al. 2001; Tiesler 2007; Walker 1995; Walker and Lucero 2000). Regarding the identification of sacrificial victims among primary human burials that lack clear signs of ancestral veneration (e.g., Fowler 1984), recent debate has been cast, specifically on the often arbitrary inferences of companion sacrifice from multiple interments. The very concept of “funerary attendants” has been questioned in many examples of alleged companion burials on the grounds of absence of direct evidence for violent death (Weiss-Krejci 2003). An alternative approach, which lends credit to the limitations in recognizing traumatic injuries in deteriorated remains, employs combined archaeological and biological profiling to distinguish patterns suggestive of ritual deposition versus reverential interment (Tiesler 2007; Tiesler and Cucina 2006). Beyond the identification of natural death versus ritual killing in retainer contexts, I wish to echo Bentley and Klaus (chapter 10) and advise caution when assigning notions of religious human sacrifice to companion burials in tomb contexts. Retainers have been discussed recurrently in this volume as one form of sacrificial immolation and have been compared to other forms of human sacrifices. On the grounds of religious theory, however such “prestigious” human killings have been denied the status of religious sacrifice, given the lack of collective

Mesoamerican Perspectives  37 7

benefits and the missing ritual etiquette that characterize most of these cases cross-culturally in the literature. A look at contact-era West Mexico illustrates this point: chroniclers describe that the death of the important paramount leader was followed by pompous public festivities, while his servants, who were thought to be in his service also in the afterworld, were slaughtered rather unceremoniously backstage by being clubbed to death (Pereira 1999). I think, therefore, that the distinct notions of “killing” those retainers—especially those confirmed to have died an unnatural death—should be explored under the broad umbrella of social theory, social status, and eschatological beliefs more than that of “human sacrifice,” given the clear diversity of documented retainer locations, forms of killing, and body-processing on the north coast of Peru. In Mesoamerica, attempts to gain a more subtle understanding of assemblages related to ancestor veneration versus other forms of ritual have proliferated, especially since the turn of the century, turning attention to the many caveats that archaeological scholarship faces in the area. At a time when the difficulties inherent in mortuary reconstruction have been admitted, cognitive and especially agency-based approaches have acquired resonance in the academic community in an attempt to imbue ritual human deposits with ideological significance and to gain a subtler understanding of the diversified ritual behaviors or attitudes substantiated in the material vestiges (Carrasco 1999; Duncan and Schwartz 2014; Kunen et al. 2002; Lucero 2003; McAnany et al. 1999; Walker 1995). In tandem with culturally contextualized social reconstructions, systematic taphonomic orientations and research styles have proliferated in different parts of Mesoamerica. The best of our regional research systematically advocates archaeothanatology (the “French School of Field Anthropology” following Duday [2009]), which has been beneficial in recording the complex ritual venues involving human sacrifice and related posthumous treatments. In the Mexico Highlands and the Maya region, the patterning of anthropogenic marks has been successfully put to work in the interpretation of ancient rituals that involved bodyprocessing (Botella et al. 2000; Chávez 2010; Pereira et al. 2006; Pijoan and Lizarraga 2004; Talavera et al. 2001; Tiesler and Cucina 2007; Turner and Turner 1999). These works have received input from a host of pre-Columbian and Colonial written sources; especially in the case of Aztec archaeology, research has found consistent links to the ritual

37 8   Tiesler

calendar, specific ceremonies, or gods (González 1985; Helfrich 1973; López Luján and Olivier 2010; Matos 1986; Moser 1973; Nájera 1987). Also, the study of native body concepts and their relationship to the Mesoamerican cosmos has contributed substantially to the understanding of the different meanings and forms of ritual immolation in Mesoamerica since the foundational work of Alfredo López Austin (1989), who formally instituted this line of research among Mesoamericanists (see also Houston et al. 2006 for the Maya area). The concepts of heavy and light body matter, body essences, and animic forces resonate well with the notion of consecration in human sacrifice and cathartic release (Hubert and Mauss 1964 [1898]). A recurrent theme in native sacrificial consumption is the sacred metamorphosis of the victim and his or her body parts through the offering of their vital essences, allowing the community to engage in the ultimate communication and exchange with the sacred. In this type of research, polisemy and epigraphy have been instrumental in understanding the multilayered notions of the body and its ethereal components. Those ethereal components were thought to reside or accumulate in specific body parts and hold vitalizaing powers: such was the case with the circulating bloody substance, “the sacred heart matter,” the “heat” of the hair on the crown of the head, and the “skull-seed.” It is revealing that the postmortem destinies and the powers retained in the mortal remains of ancestors find many analogues to those still harbored in ritually immolated bodies (although the latter typically lost their individual qualities during the act). This correlation may, for example, explain some of the similarities between reverential and postsacrificial body-processing documented among human assemblages in the Mesoamerican sphere. Similarities in placement are apparent when comparing trophy heads to those of ancestral masks, for example, or the craftsmanship of femoral relics of dynastic ancestors, in which the femoral diaphysis of sacrificial victims were turned into ritual raspa music instruments (omichicahuaztli) (Pereira 2005; Tiesler 2007). In regard to the (bio)archaeology of Andean ritual violence, it is noteworthy that some body concepts, briefly mentioned here for the Mesoamerican sphere, do indeed find parallels in the Andean world and therefore could provide useful proxies for a more in-depth exploration of body essences and their sacrificial (or natural) release in Andean deathways. Some laudable groundwork on Andean body concepts has already

Mesoamerican Perspectives  37 9

been advanced by Classen (1993) and Houston and Cummins (2004), among others. More recently, Toyne (2008:333–340; 2015b) has successfully applied emic body concepts directly to the (bio)archaeology of north coast sacrifice and the offering of vital body parts.

Concluding Remarks Beyond doubt, refinement is needed when recording and interpreting ancient human assemblages suggestive of ritual violence among ancient north coast peoples from Peru. The material and contextual characteristics, beyond aggregates of static material elements, need to be better understood according to the expected and observed ranges of sequenced ritual conduct. Within this scheme, the distinctions advocated here among modal and nonmodal mortuary behaviors, mortuary patterning, and pathway approaches hold much promise, at least on a methodological level. Beyond that, I think we need to look for a fundamental conceptual shift in interpreting ancient ritual behaviors, namely from case studies focused upon mortuary programs and their broader cultural undercurrents. This goal has been accomplished and even exceeded in this volume; the editors and contributors are to be congratulated. In this vein, the chapters unified by editors Klaus and Toyne advance a paradigm shift in the study of Andean north coast sacrifice by untangling broader social and cultural motives behind many types of ritual treatments. These differ among central and peripheral settlements, among and between the dominant, the dominated, or the conquered, and in different situations and times. This detailed level of dissecting mortuary pathways and ritual programs sets the stage for a broader discussion on the polymorph versus unified behavioral components of these killings or for inferring long-standing, culturally embedded, conservative trends as opposed to social change and crisis. Specifically, social disruption is prone to be expressed in shifts or contingencies in burial practices as seen in the material record by innovation or discontinuities in terms of mortuary repertoire. These and other conjunctions delineate new questions, agendas, and areas of research for the coming decades. For now, I thank the editors for their invitation to engage in this ambitious volume and congratulate all contributors for their substantial methodological and interpretive accomplishments.

THIS PAGE INTENTIONALLY LEFT BLANK

Reference List  3 8 1

re fe re n ce list

Abbink, Jon 2003 Love and Death of Cattle: The Paradox in Suri Attitudes Toward Livestock. Ethnos 68:341–364. Acosta, José de 2002 Natural and Moral History of the Indies. Duke University Press, Durham, North Carolina. Adams, Bradley, and John Byrd 2008 Recovery, Analysis, and Identification of Commingled Human Remains. Humana Press, New York. Adams, Bradley, and Lyle Konigsberg 2004 Estimation of the Most Likely Number of Individuals (MLNI) from Commingled Human Skeletal Remains. American Journal of Physical Anthropology 125:138–151. 2008 How Many People? Determining the Number of Individuals Represented by Commingled Human Remains. In Recovery, Analysis, and Identification of Commingled Human Remains, edited by Bradley Adams and John Byrd, pp. 241–256. Humana Press, New York. Adelson, Leslie A. 1974 The Pathology of Homicide. Charles C. Thomas, Springfield, Illinois. Agarwal, Sabrina C., and Bonnie A. Glencross 2011 Social Bioarchaeology. Wiley-Blackwell, Chichester, UK. Aldhouse-Green, Miranda J. 2001 Dying for the Gods: Human Sacrifice in Iron Age and Roman Europe. Tempus, Gloucestershire, UK. Alva, Walter 1985 Una Tumba con Máscara Funeraria de la Costa Norte del Perú. Beiträge Zur Allgemeinen Und Vergleichenden Archäologie 6:411–421. 1986 Frühe Keramik aus dem Jequetepeque-Tal, Nordperu-Cerámica temprana en el valle de Jequetepeque, norte del Perú. Materialien zur Algemeinen und Vergleichenden Archäologie. C. H. Beck, Munich. 1990 New Tomb of Royal Splendor. National Geographic 177:2–15. 1994 Sipán. Colección Cultura y Artes del Perú. General editor J. A. de Lavalle. Cervecería Backus y Johnston S.A., Lima. 2001 The Royal Tombs of Sipán: Art and Power in Moche Society. In Moche Art and Archaeology in Ancient Peru, edited by Joanne Pillsbury, pp. 223–245. National Gallery of Art, Washington, DC. 2008 Las Tumbas Reales de Sipán. In Señores de los Reinos de la Luna, edited by Krzysztof Makowski, pp. 266–279. Banco de Credito, Lima. Alva, Walter, and Susan Meneses de Alva 1984 Los Murals de Úcupe en de Valle de Zaña, Norte del Perú. Beiträge Zur Allgemeinen und Vergleichenden Archäologie 5:355–360.

3 8 2   Reference List Alva, Walter, and Luis Chero 2008 La Tumba del Sacerdote Guerrero. In Sipán: El Tesoro de la Tumbas Reales, edited by Antonio Aimi, Walter Alva, and Emilia Perassi, pp. 114–137. Giunti Industrie Grafiche, Prato, Italy. Alva, Walter, and Christopher B. Donnan 1993 Royal Tombs of Sipán. Fowler Museum of Cultural History, University of California, Los Angeles. Alva Meneses, Ignacio 2012 Ventarrón y Collud: Origin y Augue de la Civilización en la Costa Norte del Perú. Ministerio de Cultura, Lima. Ambrose, Stanley H. 1990 Preparation and Characterization of Bone and Tooth Collagen for Isotopic Analysis. Journal of Archaeological Science 17(4):431–451. Anderson-Hamilton, Laurel 2005 Cut Marks as Evidence of Pre-Columbian Human Sacrifice and Postmortem Human Bone Modification on the North Coast of Peru. Unpublished PhD dissertation, Department of Anthropology, Tulane University, New Orleans. Andrushko, Valerie A., Michele R. Buzon, Arminda M. Gibaja, Gordon F. McEwan, Antonio Simonetti, and Robert A. Creaser 2011 Investigating a Child Sacrifice Event from the Inca Heartland. Journal of Archaeological Science 38:323–333. Andrushko, Valerie A, Katie A. S. Latham, Diane Grady, Allen G. Pastron, and Phillip L. Walker 2005 Bioarchaeological Evidence for Trophy-Taking in Prehistoric Central California. American Journal of Physical Anthropology 127:375–384. Andrushko, Valerie, A., Elva C. Torres Pino, and Viviana Bellifemine 2006 The Burials at Sacsahuaman and Chokepukio: A Bioarchaeological Case Study of Imperialism from the Capital of the Inca Empire. Ñawpa Pacha 28:63–92. Arkush, Elizabeth, and Tiffiny A. Tung 2013 Patterns of War in the Andes from the Archaic to the Late Horizon: Insights from Settlement Patterns and Cranial Trauma. Journal of Archaeological Research 21:307–369. Armas, José 2008 Análisis del Material Ceramográfico de la Plaza 3C. In Investigaciones en la Huaca de la Luna 2001, edited by S. Uceda and R. Morales, pp. 177–194. Universidad Nacional de Trujillo, Trujillo. Arnold, Denise Y., and Juan de Dios Yapita 2000 El Ricón de las Cabezas. Luchas Textuales, Educación y Tierras en los Andes. ILCA and UMSA, La Paz. Arnold, Denise Y., and Christine A. Hastorf 2008 Heads of State: Icons, Power, and Politics in the Ancient and Modern Andes. Left Coast Press, Walnut Creek, California. Arriaga, Father Pablo Joseph de 1968 [1621] The Extirpation of Idolatry in Peru. Translation by L. Clark Keating. University of Kentucky Press, Louisville.

Reference List  3 8 3 Arsenault, Daniel 1988 Les Mochicas et la mort: quelques aspects idéologiques et politiques du contexte funèbre mochica. Culture 8:19–38. Asad, Talal 1993 Genealogies of Religion: Discipline and Reasons of Power in Christianity and Islam. Baltimore, Johns Hopkins University Press. Aspöck, Edeltraud 2008 What Actually Is a “Deviant Burial”? Comparing German-Language and Anglophone Research on “Deviant Burials.” In Deviant Burial in the Archaeological Record, edited by Eileen Murphy, pp. 17–34. Oxbow Books, Oxford, UK. Aufderheide, Arthur C., and Conrado Rodríguez-Martín 1998 The Cambridge Encyclopedia of Human Paleopathology. Cambridge University Press, Cambridge, UK. Baadsgaard, Aubrey, Janet Monge, and Richard L. Zettler 2012 Blugeoned, Burned, and Beautified: Reevaluating Mortuary Practices in the Royal Cemetery of Ur. In Sacred Killing: The Archaeology of Sacrifice in the Ancient Near East, edited by Anne Porter and Glenn M. Schwartz, pp. 125–158. Eisenbrauns: Winona Lake, Indiana. Bacher, Johann, Knut Wenzig, and Melanie Vogler 2005 SPSS TwoStep Cluster—A First Evaluation. Available at: http://www.statis ticalinnovations.com/products/twostep.pdf. Backo, Heather C. 2012 Distinguishing Between Natural and Cultural Taphonomic Agents in Skeletal Assemblages from Human Sacrifice Sites in Pre-Columbian Northern Coastal Peru. Unpublished PhD dissertation. Department of Anthropology, Tulane University, New Orleans. Backo, Heather C., and John W. Verano 2013 The New Temple and Temporal Continuity in Human Sacrifice at the Site of Moche, Peru. Paper presented at the 78th Annual Meeting of the Society for American Archaeology, Honolulu. Balasse, Marie, Hervé Bocherens, Andrè Mariotti, and Stanley H. Ambrose 2001 Detection of Dietary Changes by Intra-Tooth Carbon and Nitrogen Isotopic Analysis: An Experimental Study of Dentine Collagen of Cattle (Bos taurus). Journal of Archaeological Science 28:235–245. Bass, William M. 1997 Outdoor Decomposition Rates in Tennessee. In Forensic Taphonomy: The Postmortem Fate of Human Remains, edited by William D. Haglund and Marcella H. Sorg, pp. 181–186. CRC Press, Boca Raton, Florida. 2005 Human Osteology: A Laboratory and Field Manual, 4th ed. Missouri Archaeological Society, Springfield, Missouri. Bastien, Joseph W. 1978 Mountain of the Condor: Metaphor and Ritual in an Andean Ayllu. American Ethnological Society Monograph No. 64. West Publishing, St. Paul, Minnesota. Batchelor, John 1908 Ainus. In Encyclopedia of Religion and Ethics, edited by J. Hastings, pp. 239–252. Charles Scribner’s Sons, New York.

3 8 4   Reference List Bawden, Garth 1996a The Structural Paradox: Moche Culture as Political Ideology. Latin American Antiquity 6:255–273. 1996b The Moche. Blackwell, Cambridge, Massachusetts. 2001 The Symbols of Late Moche Social Transformation. In Moche Art and Archaeology in Ancient Peru, edited by Joanne Pillsbury, pp. 285–305. National Gallery of Art, Washington, DC. 2005 Ethnogenesis at Galindo, Peru. In Us and Them: Archaeology and Ethnicity in the Andes, edited by Richard Martin Reycraft, pp. 12–33. Costen Institute of Archaeology, Los Angeles. Beattie, J. H. M. 1980 Understanding Sacrifice. In Sacrifice, edited by M. F. C. Bourdillon and Meyer Fortes, pp. 29–44. Academic Press, New York. Becker, Marshall J. 1992 Burials as Caches, Caches as Burials: A New Interpretation of the Meaning of Ritual Deposits among the Classic Period Lowland Maya. In New Theories on the Ancient Maya, edited by E.  C. Danien and R.  J. Sharer, pp. 185–196. The University Museum, University of Pennsylvania, Philadelphia. 2001 Ashes to Caches: Is Dust Dust among the Heterarchical Maya? Paper presented at the 66th Annual Meeting of the Society for American Archaeology, New Orleans. Behrensmeyer, Anna K. 1975 The Taphonomy and Paleoecology of Plio-Pleistocene Vertebrate Assemblages East of Lake Rudolf, Kenya. Bulletin of the Museum of Comparative Zoology 146:473–578. Behrensmeyer, Anna K., Susan M. Kidwell, and Robert A. Gastaldo 2000 Taphonomy and Paleobiology. Paleobiology 26(sp4):103–147. Bell, Catherine 1997 Rituals: Perspectives and Dimensions. Oxford University Press, Oxford, UK Bello Galloso, Antonio 1897 [1582] Relación que Enbio a Mandar su Magestad se Hizeises desta Cuidad de Cuenca y de Todo su Provincia. Relaciónes Geográficas de Indias, Tomo III, pp. 155–196. Ministerio de Fomento, Madrid. Benfer, Robert A. 1990 The Preceramic Period Site of Paloma, Peru: Bioindications of Improving Adaptation to Sedentism. Latin American Antiquity 1:284–318. Bennett, William C. 1939 Archaeology of the North Coast of Peru: An Account of Exploration and Excavation in the Viru and Lambayeque Valleys. Anthropological Papers of the American Museum of Natural History, vol. XXXVIII, part 1. American Museum of Natural History, New York. 1950 The Gallinazo Group, Viru Valley, Peru. Yale University Press, New Haven. Benson, Elizabeth P. 1972 The Mochica: A Culture of Peru. Praeger, New York. 1995 Art, Agriculture, Warfare and the Guano Islands. In Andean Art: Visual

Reference List  3 8 5 Expression and Its Relation to Andean Beliefs and Values, edited by Penelope Z. Dransart, pp. 245–264. Avebury, Aldershot, UK. 2001 Why Sacrifice? In Ritual Sacrifice in Ancient Peru, edited by Elizabeth P. Benson and Anita G. Cook, pp. 1–20. University of Texas Press, Austin. 2012 The Worlds of The Moche on the North Coast of Peru. University of Texas Press, Austin. Benson, Elizabeth P., and Anita G. Cook (editors) 2001 Ritual Sacrifice in Ancient Peru. University of Texas Press, Austin. Beresford-Jones, David G., Susana Arce T, Oliver Q. Whaley, and Alex J. Chepstow-Lusty 2009 The Role of Prosopis in Ecological and Landscape Change in the Samaca Basin, Lower Ica Valley, South Coast Peru from the Early Horizon to the Late Intermediate Period. Latin American Antiquity 20:303–332. Bernuy, Katiusha, and Vanessa Bernal 2008 La Tradición Cajamarca en San José de Moro: Una Evidencia de Interacción Interregional Durante el Horizonte Medio. In Arqueología Mochica: Nuevos Enfoques, edited by Luis Jaime Castillo, Helaine Bernier, Julio Rucabado, and Gregory Lockard, pp. 67–80. Fondo Editorial de la Pontificia Universidad Católica del Perú, Lima. Berrin, Kathleen (editor) 1997 The Spirit of Ancient Peru: Treasures from the Museo Arqueológico Rafael Larco Herrera. Fine Arts Museums of San Francisco, San Francisco. Berryman, Hugh E., and Susan Haun 1996 Applying Forensic Techniques to Interpret Cranial Fracture Patterns in an Archaeological Specimen. International Journal of Osteoarchaeology 6:2–9. Berryman, Hugh E., and Steven A. Symes 1998 Recognizing Gunshot and Blunt Cranial Trauma through Fracture Interpretation. In Forensic Osteology, edited by Kathleen J. Reichs, pp. 333–352. Charles Thomas, Springfield, Illinois. Besom, Thomas 2009 Of Summits and Sacrifice: An Ethnohistoric Study of Inka Religious Practices. University of Texas Press, Austin. Betanzos, Juan de 1996 [1557] Narrative of the Incas. Translation by Roland Hamilton and Dana Buchanan. University of Texas Press, Austin. Bethard, Jonathan D., Catherine Gaither, Víctor Vásquez Sánchez, Teresa Rosales Tham, and Jonathan D. Kent 2008 Isótopos Estables, Dieta y Movilidad de los Pobladores de un Conjunto Residencial en Santa Rita B, Valle de Chao, Perú. Revista de Bioarqueología Archaeobios 2:19–27. Bezúr, Anikó 2003 Variability in Sicán Copper Alloy Artifacts. Unpublished PhD dissertation. Department of Materials Science and Engineering, University of Arizona, Tucson. 2014 Tecnología y Organización de la Producción de Cobre Arsenical Sicán.

3 8 6   Reference List In Cultura Sicán: Esplendor Preinca de la Costa Norte, edited by Izumi Shimada. Fondo Editorial del Congreso de Perú, Lima. Bhabba, Homi K. 1994 The Location of Culture. Routledge, London. Billman, Brian 1997 Population Pressure and the Origins of Warfare in the Moche Valley, Peru. In Integrating Archaeological Demography: Multidisciplinary Approaches to Prehistoric Populations, edited by Richard R. Paine, pp. 285–309. Center for Archaeological Investigations, Occasional Paper 24. Southern Illinois University, Carbondale. 2002 Irrigation and the Origins of the Southern Moche State on the North Coast of Peru. Latin American Antiquity 13:371–400. Binford, Lewis 1971 Mortuary Practices: Their Study and Their Potential. In Approaches to the Social Dimension of Mortuary Practices, edited by James Brown, pp. 6–29. Memoirs of the Society for American Archaeology, No. 25. 1981 Bones: Ancient Men and Modern Myths. Academic Press, New York. Bird-David, Nurit 1990 The Giving Environment: Another Perspective on the Economic System of Gatherer-Hunters. Current Anthropology 1:189–196. Bishop, Neil A., and Christopher J. Knüsel 2005 A Palaeodemographic Investigation of Warfare in Prehistory. In Warfare, Violence, and Slavery in Prehistory, edited by Michael Parker Pearson and I.  N.  J. Thorpe, pp. 201–216. British Archaeological Research International Series 1374, Oxford, UK. Blackwood, Peter 1981 Rank, Exchange, and Leadership in Four Vanuatu Societies. In Vanuatu: Politics, Economics, and Ritual in Island Melanesia, edited by M. R. Allen, pp. 35–84. Academic Press, New York. Bloch, Maurice 1992 Prey into Hunter: The Politics of Religious Experience. Cambridge University Press, Cambridge, UK. Bloch, Maurice, and Jonathan Parry 1982 Introduction: Death and the Regeneration of Life. In Death and the Regeneration of Life, edited by Maurice Bloch and Jonathan Parry, pp. 1–44. Cambridge University Press, Cambridge, UK. Blom, Deborah E. 2005 Embodying Borders: Human Body Modification and Diversity in Tiwanaku Society. Journal of Anthropological Archaeology 24:1–24. Blom, Deborah E., Jane E. Buikstra, Linda Keng, Paula D. Tomczak, Eleanor Shoreman, and Debbie Stevens-Tuttle 2005 Anemia and Childhood Mortality: Latitudinal Patterning Along the Coast of Pre-Columbian Peru. American Journal of Physical Anthropology 127:152–169. Blom, Deborah E., and John Wayne Janusek 2004 Making Place: Humans as Dedications in Tiwanaku. World Archaeology 36(1):123–141.

Reference List  3 87 Blom, Deborah E., John Wayne Janusek, and Jane E. Buikstra 2003 A Re-Evaluation of Human Remains from Tiwanaku. In Tiwanaku and Its Hinterlands: Archaeology and Paleoecology of an Andean Civilization, edited by Alan Kolata, pp. 435–446. Vol. 2. Smithsonian Institution Press, Washington, DC. Boaz, Neil T., and Anna K. Behrensmeyer 1976 Hominid Taphonomy: Transport of Human Skeletal Parts in an Artificial Fluvial Environment. American Journal of Physical Anthropology 45:53–60. Bolin, Inge 1998 Rituals of Respect: The Secret of Survival in the High Peruvian Andes. University of Texas Press, Austin. Bonavia, Duccio 1985 Mural Painting in Ancient Peru. Translation by Patricia J. Lyon. Indiana University Press, Bloomington. 2008 The South American Camelids. Translation by J. F. Espinoza. Cotsen Institute of Archaeology, University of California, Los Angeles. Bonnier, Elisabeth 1997 Preceramic Architecture in the Andes: The Mito Tradition. In Arqueologia Peruana 2: Arquitectura y Civilización en los Andes Prehispánicos, edited by Elisabeth Bonnier and Henning Bischoff, pp. 120–144. Sociedad Arqueológica Peruano-Alemana, Reiss-Museum, Manheim. Boone, Elizabeth Hill 1984 Ritual Human Sacrifice in Mesoamerica. Dumbarton Oaks, Washington, DC. 2000 Stories in Red and Black: Pictorial Histories of the Aztecs and Mixtecs. 1st ed. University of Texas Press, Austin. Botella, Miguel C., Inmaculada Alemán, and Sylvia A. Jiménez 2000 Los Huesos Humanos. Manipulación y Alteraciones. Bellaterra, Barcelona. Bourdillon, Michael F. C. 1980 Introduction. In Sacrifice, edited by Michael F. C. Bourdillon and Meyer Fortes, pp. 1–27, Academic Press, London. Bourdillon, M. F. C., and Meyer Fortes (editors) 1980 Sacrifice. Academic Press, New York. Bourget, Steve 1994 Bestiaire Sacré et Flore Magique: Écologie Rituelle de l’iconographie de la Culture Mochica, Côte Nord du Pérou. Unpublished Ph.D. dissertation, Département d’anthropologie, Université de Montréal, Québec. 1997 Las Excavaciones en la Plaza 3A de la Huaca de la Luna. In Investigaciones en la Huaca de la Luna 1995, edited by Santiago Uceda, Elìas Mujica, and Ricardo Morales, pp. 51–59. Facultad de Ciensas Sociales, Universidad Nacional de la Libertad, Trujillo. 1998a Pratiques Sacrificielles and Funéraires au site Moche de la Huaca de la Luna, Côte Nord du Pérou. Bulletin de l’Institut Français d’Études Andines 27:41–74. 1998b Excavaciones en la Plaza 3A y en la Platforma II de la Huaca de la Luna Durante 1996. In Investigaciones en la Huaca de la Luna 1996, edited by Santiago Uceda, Elìas Mujica, and Ricardo Morales, pp. 43–64. Facultad de Ciencias Sociales, Universidad Nacional de la Libertad, Trujillo.

3 8 8   Reference List 2001a Children and the Ancestors: Ritual Practices at the Moche Site of Huaca de la Luna, North Coast of Peru. In Ritual Sacrifice in Ancient Peru, edited by Elizabeth P. Benson and Anita G. Cook, pp. 93–118. University of Texas Press, Austin. 2001b Rituals of Sacrifice: Its Practice at Huaca de la Luna and Its Representation in Moche Iconography. In Moche Art and Archaeology in Ancient Peru, edited by Joanne Pillsbury, pp. 89–109. National Gallery of Art, Washington, DC. 2003 Somos Diferentes: Dinámica Ocupacional del Sitio Castillo de Huancaco, Valle de Virú. In Moche: Hacia el final del milenio; Actas del segundo Coloquio sobre la Cultura Moche, Trujillo, 1 al 7 de agosto de 1999, edited by Santiago Uceda and Elias Mujica, pp. 245–267. Universidad Nacional de Trujillo, Peru, and Fondo Editorial de la Pontificia Universidad Católica del Perú, Lima. 2005 Who Were the Priests, the Warriors, and the Prisoners? A Peculiar Problem of Identity in Moche Culture and Iconography, North Coast of Peru. In Us and Them: Archaeology and Ethnicity in the Andes, edited by Richard Martin Reycraft, pp. 73–85. Cotsen Institute of Archaeology Press, University of California, Los Angeles. 2006 Sex, Death, and Sacrifice in Moche Religion and Visual Culture. University of Texas Press, Austin. 2010 Cultural Assignations During the Early Intermediate Period: The Case of Huancaco, Virú Valley. In New Perspectives on Moche Political Organization, edited by Jeffrey Quilter and Luis Jaime Castillo, pp. 201–222. Dumbarton Oaks, Washington, DC. Bourget, Steve, Bruno Alva, and Kimberly L. Jones 2012 El Señor de Úcupe: La Tumba de un Noble Mochica en la Huaca El Pueblo, Valle de Saña. In Tesoros Preincas de la Cultura Mochica, edited by Luis Hurtado, pp. 99–109. Fundación Wiese, Lima. Bourget, Steve, and Kimberly L. Jones (editors) 2008 The Art and Archaeology of the Moche: An Ancient Andean Society of the Peruvian North Coast. University of Texas Press, Austin. Bourget, Steve, and Jean-François Millaire 2000 Excavaciones en la Plaza 3A y Plataforma II de la Huaca de la Luna. In Investigaciones en la Huaca de la Luna 1997, edited by Santiago Uceda, Elías Mujica, and Ricardo Morales, pp. 47–60. Facultad de Ciencias Sociales, Universidad Nacional de La Libertad, Trujillo. Bourget, Steve, and Margaret Newman 1998 A Toast to the Ancestors: Ritual Warfare and Sacrificial Blood in the Moche Culture. Baessler-Archiv 46:85–106. Brickley, Megan, and Martin Smith 2006 Culturally Determined Patterns of Violence: Biological Anthropological Investigations at a Historic Urban Cemetery. American Anthropologist 108:163–177. Bromage, Timothy G. 1987 The Scanning Electron Microscopy/Replica Technique and Recent

Reference List  3 8 9 Applications to the Study of Fossil Bone. Scanning Microscopy 1:607– 613. Bromage, Timothy G., and Alan Boyde 1984 Microscopic Criteria for the Determination of Directionality of Cutmarks on Bone. American Journal of Physical Anthropology 65:359–366. Brooks, William E., Matthew G. Brooks, Jonathan D. Kent, Victor F. Vásquez S., Teresa E. Rosales T., Rosenelsy Marrero Cuebas, and Jason C. Willett 2001 The Muralla Pircada-Evidence for a 2000–Year-Old Natural Hazards Dam Near the Santa Rita B Archaeological Site, Northern Perú. Geological Society of America Abstracts with Programs 33, n. 6:A-128. Brooks, William E., Jason C. Willett, Jonathan D. Kent, Victor Vasquez, and Teresa Rosales 2005 The Muralla Pircada—An Ancient Andean Debris Flow Retention Dam, Santa Rita B Archaeological Site, Chao Valley, Northern Perú. Landslides 2:117–123. Browman, David L. 1990 Camelid Pastoralism in the Andes: Llama Caravan Fleteros, and Their Importance in Production and Distribution. In Nomads in a Changing World, edited by Phillip Carl Salzman and John G. Galaty, pp. 395–438. Instituto Universitario Orientale, Naples, Italy. Brown, James A. 2010 Cosmological Layouts of Secondary Burials as Political Instruments. In Mississippian Mortuary Practices: Beyond Hierarchy and the Representationist Perspective, edited by Lynne P. Sullivan and Robert Mainfort, pp. 30–52. University Press of Florida, Gainesville. Brown, Michael F. 1984 The Role of Words in Aguaruna Hunting Magic. American Ethnologist 11:545–558. Browne, David M., Helaine Silverman, and Rubén García 1993 A Cache of 48 Nasca Trophy Heads from Cerro Carapo, Peru. Latin American Antiquity 4:274–294. Brüning, Hans Heinrich 1922 Lambayeque: Estudios Monográficos. Liberería e Imprenta de Dionisio Mendoza, Chiclayo, Perú. Buikstra, Jane E. 1977 Biocultural Dimensions of Archeological Study: A Regional Perspective. In Biocultural Adaptation in Prehistoric America, edited by Robert L. Blakely, pp. 67–84. Southern Anthropological Society Proceedings No. 11. University of Georgia Press, Athens. Buikstra, Jane E., and Douglas H. Ubelaker (editors) 1994 Standards for Data Collection from Human Skeletal Remains. Arkansas Archeological Survey Research Series No. 44, Fayetteville, Arkansas. Bulliet, Richard W. 2005 Hunters, Herders, and Hamburgers: The Past and Future of Human-Animal Relationships. Columbia University Press, New York.

3 9 0  Reference List Burger, Richard L. 1992a Chavin and the Origins of Andean Civilization. Thames and Hudson, New York. 1992b The Sacred Center of Chavín de Huantar. In The Ancient Americas: Art from Sacred Landscapes, edited by Richard F. Townsend, pp. 265–277. Art Institute of Chicago, Chicago. Bush, Helen 1991 Concepts of Health and Stress. In Health in Past Societies. Biocultural Interpretations of Human Skeletal Remains in Archaeological Contexts, edited by Helen Bush and Marek Zvelebil, pp. 11–22. BAR International Series 567, Oxford, UK. Bush, Helen, and Ann Stirland 1991 Romano-British Decapitation Burials: A Comparison of Osteological Evidence and Burial Ritual from Two Cemeteries. Anthropologie (Brno) 29:205–210. Bush, Mark B., Barbara C. S. Hansen, Donald T. Rodbell, Geoffrey O. Seltzer, Kenneth R. Young, Blanca León, Mark B. Abbott, Miles R. Silman, and William D. Gosling 2005 A 17,000-year History of Andean Climate and Vegetation Change from Laguna de Chochos, Peru. Journal of Quaternary Science 20 (7–8):703–714. Buzon, Michele R., and Margaret A. Judd 2008 Investigating Health at Kerma: Sacrificial Versus Nonsacrificial Individuals. American Journal of Physical Anthropology 136:93–99. Byrd, Jason H., and James L. Castner (editors) 2001 Forensic Entomology: The Utility of Arthropods in Forensic Investigations. CRC Press, Boca Raton, Florida. Cabello Balboa, Miguel 1951 [1586] Miscelánea Antártica: Una Historia del Perú Antiguo. Lima: Instituto de Ethnología, Universidad Nacional Mayor de San Marcos. Calancha, Antonio de 1972 [1638] Crónica Moralizada del Orden de San Agustin en el Perú, con Sucesos Esemplares en esta Monarquía. Pedro Lacavalleria, Barcelona. Campbell, Roderick 2007 Flesh, Blood, and Bones: Kinship and Violence in the Social Economy of the Late Shang. Unpublished PhD dissertation, Department of Anthropology, Harvard University, Cambridge, Massachusetts. 2012 On Sacrifice: An Archaeology of Shang Sacrifice. In Sacred Killing: The Archaeology of Sacrifice in the Ancient Near East, edited by Anne Porter and Glenn M. Schwartz, pp. 305–323. Eisenbrauns, Winona Lake, Indiana. Campobasso, C.P. 2001 Factors Affecting Decomposition and Diptera Colonization. Forensic Science International 120(1–2):18–27. Canti, M.G. 2003 Earthworm Activity and Archaeological Stratigraphy: A Review of Products and Processes. Journal of Archaeological Science 30:135–148.

Reference List  3 9 1 Carneiro, Robert L. 1970 Hunting and Hunting Magic among the Amahuaca of the Peruvian Montaña. Ethnology 9:331–341. Carrasco, David 1999 City of Sacrifice: The Aztec Empire and the Role of Violence in Civilization. Beacon, Boston. Carrion, Rebecca 1940 La Luna y su Personificación Ornitomórfa en el Arte Chimú. Actas y Trabajos Científicos de 27 Congreso Internacional del Americanistas 1:571–587. Carson, E. A., Vincent Stefan, and Joseph Powell 2000 Skeletal Manifestations of Bear Scavenging. Journal of Forensic Sciences 45:515–526. Carter, Elizabeth 2012 On Human and Animal Sacrifice in the Late Neolithic at Domuztepe. In Sacred Killing: The Archaeology of Sacrifice in the Ancient Near East, edited by Anne Porter and Glenn M. Schwartz, pp. 97–124. Eisenbrauns, Winona Lake, Indiana. Castillo, Luis Jaime 2000a La Ceremonia del Sacrificio: Batallas y Muerte en el Arte Mochica. AFP Integra, Museo Arqueológico Rafael Larco Herrera, Lima. 2000b La presencia de Wari en San José de Moro. In Los Dioses del Antiguo Perú, edited by Krzysztof Makowski, pp. 103–135. Colección Arte y Tesoros del Perú. Banco de Crédito del Perú, Lima. 2001 The Last of the Mochicas: A View from the Jequetepeque Valley. In Moche Art and Archaeology in Ancient Peru, edited by Joanne Pillsbury, pp. 307– 332. National Gallery of Art, Washington, DC. 2003 Los Últimos Mochicas en Jequetepeque. In Moche Hacia el Final Del Milenio, Tomo II, edited by Santiago Uceda and Elías Mujica, pp. 65–123. Universidad Nacional de Trujillo y Pontificia Universidad Católica del Perú, Lima. 2005 Las Señoras de San José de Moro, Rituales funerarios en la costa norte del Perú, In Divina y Humana: La Mujer en los Antiguos México y Perú. Instituto Nacional de Antropología e Historia, Conaculta, Mexico. 2010 Moche Politics in the Jequetepeque Valley: A Case for Political Opportunism. In New Perspectives on Moche Political Organization, edited by Jeffrey Quilter and Luis Jaime Castillo, pp. 83–109. Dumbarton Oaks, Washington, DC. 2011 San José de Moro y la Arqueología de Valle de Jequetepeque. Fondo Editoral de Pontificia Universidad Católica del Perú, Lima. 2012 San José de Moro y el Fin de los Mochicas en el Valle de Jequetepeque, Costa Norte del Perú. Unpublished PhD dissertation, Department of Anthropology, University of California, Los Angeles. Castillo, Luis Jaime, and Christopher B. Donnan 1994a Los Mochicas del Norte y los Mochicas del Sur: Una Perspectiva Desde el Valle de Jequetepeque. In Vicús, edited by Krzysztof Makowski, pp. 143– 181. Colección Arte y Tesoros del Perú. Banco de Crédito del Perú, Lima.

3 9 2   Reference List 1994b La Ocupación Moche de San José de Moro, Jequetepeque. In Moche: Propuestas y Perspectivas, edited by Santiago Uceda and Elías Mujica, pp. 93–146. Travaux de l’Institut Français d’Etudes Andines, Lima. Castillo, Luis Jaime, and Ulla Holmquist 2000 La Ceremonia del Sacrificio: Batallas y Muerte en el Arte Mochica. In La Ceremonia del Sacrificio: Batallas y Muerte en el Arte Mochica (exhibition catalog), pp. 14–29. Museo Arqueológico Rafael Larco Herrera, Lima. Castillo, Luis Jaime, and Jeffrey Quilter 2010 Many Moche Models. In New Perspectives on Moche Political Organization, edited by Jeffrey Quilter and Luis Jaime Castillo, pp. 1–16. Dumbarton Oaks, Washington, DC. Castillo, Luis Jaime, and Carlos Rengifo 2008a Identidades Funerarias Femeninas y Poder Ideológico en las Sociedades Mochicas. In Los Señores de los Reinos de la Luna, edited by Krzysztof Makowski, pp. 2–34. Colección de Arte y Tesoros del Perú, Banco de Crédito del Perú, Lima. 2008b El Género y el Poder: San José de Moro. In Los Señores de los Reinos de la Luna, edited by Krzysztof Makowski, pp. 165–182, Colección Arte y Tesoros del Perú, Banco de Crédito del Perú, Lima. Castillo, Luis Jaime, Julio Rucabado, Martín del Carpio, Katiusha Bernuy, Karim Ruiz, Carlos Rengifo, Gabriel Prieto, and Carole Fraresso 2008 Ideología y Poder en la Consolidación, Colapso y Reconstitución del Estado Mochica del Jequetepeque: El Proyecto Arqueológico San José de Moro (1991–2006). Ñawpa Pacha 29:1–86. Castillo, Luis Jaime, and Santiago Uceda 2008 The Mochicas. In Handbook of South American Archaeology, edited by Helaine Silverman and William Isbell, pp. 707–729. Springer, New York. Cavallaro, Rafael 1997 Architectural Analysis and Dual Organization in the Andes. In Architecture and Civilization in the Prehispanic Andes, edited by Elisabeth Bonnier and Henning Bischof, pp. 120–144. Sociedad Arqueológica PeruanoAlemana, Reiss-Museum, Mannheim. Centurión, Jorge, and Manuel Curo 2003 Proyecto Evaluación Arqueológico de Cerro Cerrillos. Technical report on file at the Museo Nacional de Arqueología y Etnografía Hans Henrich Brüning, Lambayeque, Perú. Ceruti, Maria Constanza 2003 Llullaillaco: Sacrificios y Ofrendas en un Santuario Inca de Alta Montaña. Ediciones Universidad Católica de Salta, Salta, Argentina. 2004 Human Bodies as Objects of Dedication at Inca Mountain Shrines (North-Western Argentina). World Archaeology 36:103–122. Cervantes, Gabriela, Izumi Shimada, Haagen D. Klaus, Kelly J. Knudson, and Ken-ichi Shinoda 2011 Multi-Ethnicity in the Sicán World: Figurines and Other Lines of Evidence. Paper presented at the 76th Annual Meeting of the Society for American Archaeology, Sacramento, California.

Reference List  3 9 3 Chamberlain, Andrew, and Mike Parker Pearson 2001 Earthly Remains: The History and Science of Preserved Human Bodies. Oxford University Press, Oxford, UK. Chambers, Sarah C. 1999 “To the Company of a Man Like My Husband, No Law Can Compel Me”: The Limits of Sanctions against Wife Beating in Arequipa, Peru, 1780– 1850. Journal of Women’s History 11:31–52. Chapdelaine, Claude 1998 Excavaciones en la Zona Urbana de Moche Durante 1996. In Investigaciones en la Huaca de la Luna 1996, edited by Santiago Uceda, Elías Mujica and Ricardo Morales, pp. 85–115. Facultad de Ciencias Sociales, Universidad Nacional de La Libertad, Trujillo. 2000 Struggling for Survival: The Urban Class of the Moche Site, North Coast of Peru. In Environmental Disaster and the Archaeology of Human Response, edited by Garth Bawden and Richard Martin Reycraft, pp. 121–142. Maxwell Museum of Anthropology, Anthropological Papers 7. University of New Mexico Press, Albuquerque. 2001 The Growing Power of a Moche Urban Class. In Moche Art and Archaeology in Ancient Peru, edited by Joanne Pillsbury, pp. 69–87. National Gallery of Art, Washington, DC. 2008 Moche Art Style in the Santa Valley: Between Being “à la Mode” and Developing a Provincial Identity. In The Art and Archaeology of the Moche: An Ancient Andean Society of the Peruvian North Coast, edited by Steve Bourget and Kimberly L. Jones, pp. 129–152. University of Texas Press, Austin. 2010 Moche Political Organization in the Santa Valley: A Case of Direct Rule through Gradual Control of the Local Population. In New Perspectives on Moche Political Organization, edited by Jeffrey Quilter and Luis Jaime Castillo, pp. 252–279. Dumbarton Oaks, Washington, DC. 2011a Los Moches del Santa: Una Larga Historia. In Arqueologia de la Costa de Ancash, edited by Miłosz Giersz and Iván Ghezzi, pp. 185–230. ANDES Boletín del Centro de Estudios Precolombinos de la Universidad de Varsovia 8. Centro de Estudios Precolombinos de la Universidad de Varsovia/Institut Français d’Études Andines, Warsaw and Lima. 2011b Recent Advances in Moche Archaeology. Journal of Archaeological Research 19:191–231. Chapdelaine, Claude, Víctor Pimentel, and Jorge Gamboa 2005 Contextos Funerarios Moche del Sitio El Castillo de Santa: Una Primera Aproximación. Corriente Arqueológica 1:13–41. Chávez, Ximena 2010 Rituales Funerarios en el Templo Mayor de Tenochtitlán. Instituto Nacional de Antropología e Historia, Mexico City. Chero, Luis 2012 Los Guardianes de las Tumbas Reales de Sipán. In Tesoros Preincas de la Cultura Mochica, edited by Luis Hurtado, pp. 35–43. Fundación Wiese, Lima. Chicoine, David 2004 The Moche Presence in the Nepeña Valley: View from Huambacho.

3 9 4   Reference List Paper presented at the 64th Annual Meeting of the Society for American Archaeology, Montréal, Québec. 2006 Early Horizon Architecture at Huambacho, Nepeña Valley, Peru. Journal of Field Archaeology 31:1–22. 2011 Death and Religion in the Southern Moche Periphery: Funerary Practices at Huambacho, Nepeña Valley, Peru. Latin American Antiquity 22:525–548. Christman, Jared 2008 The Gilgamesh Complex: The Quest for Death Transcendence and the Killing of Animals. Society and Animals 16:297–315. Cieza de León, Pedro de 1963 [1538] The Second Part of the Chronicle of Peru. Translation by Clements R. Markham. Burt Franklin, New York. 1967 [1538] El Señorio de los Incas (Segundo Parte de La Crónica del Perú). Instituto de Estudios Peruanos, Lima. 1998 [1533] The Discovery and Conquest of Peru, translated and edited by Alexandra Parma Cook and Nobel David Cook. Duke University Press, Durham, North Carolina. Classen, Constance 1993 Inca Cosmology and the Human Body. University of Utah Press, Salt Lake City. Cleland, Kathryn M., and Izumi Shimada 1992 Sicán Bottles: Marking Time in the Peruvian Bronze Age. Andean Past 3:193–235. 1998 Paleteada Potters: Technology, Production Sphere, and Sub-Culture in Ancient Peru. In Andean Ceramics: Technology, Organization, and Approaches, edited by Izumi Shimada, 111–150. Museum Applied Science Center for Archaeology and University of Pennsylvania Museum of Archaeology and Anthropology, Philadelphia. Coale, Ansley, and Paul Demeny 1966 Regional Model Life Tables and Stable Populations. Princeton University Press, Princeton, New Jersey. Cobo, Bernabe 1990 [1653] Inca Religion and Customs. Translation by Roland Hamilton. University of Texas Press, Austin. Coe, William R. 1959 Piedras Negras Archaeology: Artifacts, Caches, and Burials. University Museum monographs. University of Pennsylvania, Philadelphia. 1965 Caches and Offertory Practices of the Maya Lowlands. In Handbook of Middle American Indians, Vol. 2, edited by R. Wauchope, pp. 462–468. University of Texas Press, Austin. Cohen, Mark, and Sharon Bennett 1993 Skeletal Evidence for Sex Roles and Gender Hierarchies in Prehistory. In Sex Roles and Gender Hierarchies, edited by B. Miller, pp. 93–97. Cambridge University Press, Cambridge, UK. Cole, Jennifer 1997 Sacrifice, Narratives, and Experience in East Madagascar. Journal of Religion in Africa 27:401–425.

Reference List  3 9 5 Conrad, Geoffrey W. 1982 The Burial Platforms of Chan Chan: Some Social and Political Implications. In Chan Chan: Andean Desert City, edited by Michael Moseley and Kent Day, pp. 87–117. University of New Mexico Press, Albuquerque. Cook, Anita G. 1987 The Middle Horizon Ceramic Offerings from Conchopata. Ñawpa Pacha 22:49–90. 2001 Huari D-Shaped Structures, Sacrificial Offerings, and Divine Rulership. In Ritual Sacrifice in Ancient Peru, edited by Elizabeth P. Benson and Anita G. Cook, pp. 137–163. University of Texas Press, Austin. Cordy-Collins, Alana 1992 Archaism or Tradition? The Decapitation Theme in Cupisnique and Moche Iconography. Latin American Antiquity 3(3):206–220. 1996a Lambayeque. In Andean Art at Dumbarton Oaks, Volume I, edited by Elizabeth Hill Boone, pp. 189–222. Dumbarton Oaks, Washington, DC. 1996b Chimu. In Andean Art at Dumbarton Oaks, Volume I, edited by Elizabeth Hill Boone, pp. 223–276. Dumbarton Oaks, Washington, DC. 2001a Decapitation in Cupisnique and Early Moche Societies. In Ritual Sacrifice in Ancient Peru, edited by Elizabeth P. Benson and Anita G. Cook, pp. 21–33. University of Texas Press, Austin. 2001b Blood and the Moon Priestess: Spondylus Shells in Moche Ceremony. In Ritual Sacrifice in Ancient Peru, edited by Elizabeth P. Benson and Anita G. Cook, pp. 35–53. University of Texas Press, Austin. 2009 Curated Corpses and Scattered Skeletons: New Data and New Questions Concerning the Moche Dead. In Andean Civilization: A Tribute to Michael Moseley, edited by Joyce Marcus and Patrick Ryan Williams, pp. 181–194. Cotsen Institute of Archaeology Press, Los Angeles. Cordy-Collins, Alana, and Charles F. Merbs 2003 Forensic Iconography and the Moche Giants. Paper presented at the Symposium: The Art, the Arts, and the Archaeology of the Moche. Paper presented at the 4th D. J. Sibley Family Conference on World Traditions of Culture and Art, Austin, Texas. Corruccini, Robert S., and Izumi Shimada 2002 Dentally Determined Biological Kinship in Relation to Mortuary Patterning of Human Remains from Huaca Loro, Peru. American Journal of Physical Anthropology 117:113–121. Coughlan, Jennifer, and Malin Holst 2000 Health Status. In Blood Red Roses: The Archaeology of a Mass Grave from the Battle of Towton AD 1461, edited by Veronica Fiorato, Anthea Boylston, and Christopher Knüsel, pp. 60–76. Oxbow, Cambridge, UK. Counts, Dorothy A. Judith K. Brown, and Jacquelyn C. Campbell 1999 To Have and to Hit: Cultural Perspectives on Wife Beating. University of Illinois Press, Urbana. Craig, Alan K. 1985 Cis-Andean Environmental Transects: Late Quaternary Ecology of Northern and Southern Peru. In Andean Ecology and Civilization, edited by

3 9 6   Reference List Shozo Masuda, Izumi Shimada, and Craig Morris, pp. 22–44. University of Tokyo Press, Tokyo. Craig, Alan K., and Izumi Shimada 1986 El Niño Flood Deposits at Batán Grande, Northern Peru. Geoarchaeology 1:29–38. Craine, Joseph M., Andrew J. Elmore, Marcos P. M. Aidar, Mercedes Bustamante, Todd E. Dawson, Erik A. Hobbie, Ansgar Kahmen, Michelle C. Mack, Kendra K. McLauchlan, Anders Michelsen, Gabriela B. Nardoto, Linda H. Pardo, Josep Peñuelas, Peter B. Reich, Edward A. G. Schuur, William D. Stock, Pamela H. Templer, Ross A. Virginia, Jeffrey M. Welker, and Ian J. Wright 2009 Global Patterns of Foliar Nitrogen Isotopes and Their Relationships with Climate, Mycorrhizal Fungi, Foliar Nutrient Concentrations, and Nitrogen availability. New Phytologist 183:980–992. Crist, Thomas A.J., Arthur Washburn, Hydrow Park, Ian Hood, and Molly A. Hickey 1997 Cranial Bone Displacement as a Taphonomic Process in Potential Child Abuse Cases. In Forensic Taphonomy: The Postmortem Fate of Human Remains, edited by William D. Haglund and Marcella H. Sorg, pp. 319– 336. CRC Press, Boca Raton, Florida. Croft, Paul 2003 The Animal Bones. In The Colonisation and Settlement of Cyprus: Investigations at Kissonerga-Mylouthkia, 1976–1996, edited by Edgar J. Peltenburg and D. L. Bolger, pp. 49–58. Paul Åströms, Göteborg. Cross, Pamela J. 2011 Horse Burial in First Millennium AD Britain: Issues of Interpretation. European Journal of Archaeology 14:190–209. Cucina, Andrea, and Vera Tiesler 2007 Nutrition, Lifestyle, and Social Status of Skeletal Remains from Nonfunerary and “Problematical” Contexts. In New Perspectives on Human Sacrifice and Ritual Body Treatment in Ancient Maya Society, edited by Vera Tiesler and Andrea Cucina, pp. 251–262. Springer, New York. Currey, James D., and G. Butler 1975 The Mechanical Properties of Bone Tissue in Children. Journal of Bone and Joint Surgery, 57–A, 810–814. Cutright, Robin E. 2009 Between the Kitchen and the State: Domestic Practice and Chimu Expansion in the Jequetepeque Valley, Peru, Unpublished PhD dissertation, Department of Anthropology, University of Pittsburgh, Pennsylvania. D’Altroy, Terence N. 2002 The Incas. Blackwell, Malden, Massachusetts. Danforth, Marie Elaine 1999 Nutrition and Politics in Prehistory. Annual Review of Anthropology 28:1–25. Day, Kent 1982 Ciudadelas: Their Form and Function. In Chan Chan: Andean Desert City, edited by Michael Moseley and Kent Day, pp. 55–66. University of New Mexico Press, Albuquerque.

Reference List  3 97 de Herrera, Antonio 1944 Historia General de los Hechos de los Castellanos, 10 vols. Asunción, Paraguay: Editorial Guariana. de Sales, Anne 1980 Deux Conceptions de l’alliance à travers la fête de l’ours en Sibérie. Études Mongoles 11:147–213. de Xérez, Francisco 1872 [1534] Reports on the Discovery of Peru. Translation by Clements R. Markham. Hakluyt Society, London, England. del Angel, Andres, and Hector B. Cisneros 2004 Technical Note: Modification of Regression Equations Used to Estimate Stature in Mesoamerican Skeletal Remains. American Journal of Physical Anthropology 125:264–265. del Carpio Perla, Martín 2008 La Ocupación Mochica Medio en San José de Moro. In Arqueología Mochica: Nuevos Enfoques, edited by Luis Jaime Castillo, Helaine Bernier, Julio Rucabado and Gregory Lockard, pp. 81–104. Fondo Editorial de la Pontificia Universidad Católica del Perú, Lima. deFrance, Susan D. 2010 Paleopathology and Health of Native and Introduced Animals on Southern Peruvian and Bolivian Spanish Colonial Sites. International Journal of Osteoarchaeology 20:508–524. Delibes, Rocío, and Alfonso Barragán 2008 Consumo Ritual de Chicha en San José de Moro. In Arqueología Mochica: Nuevos Enfoques, edited by Luis Jaime Castillo, Helene Bernier, Gregory Lockard and Julio Rucabado, pp. 105–118. Fondo Editorial de la Pontificia Universidad Católica del Perú, Lima. DeNiro, Michael J. 1985 Postmortem Preservation and Alteration of In Vivo Bone Collagen Isotope Ratios in Relation to Palaeodietary Reconstruction. Nature 317:806–809. 1988 Marine Food Sources for Prehistoric Coastal Peruvian Camelids: Isotopic Evidence and Implications. In Economic Prehistory of the Central Andes, edited by Elisabeth S. Wing and Jane C. Wheeler, pp. 119–128. British Archaeological Reports, International Series 427. Archaeopress, Oxford, UK. DeNiro, Michael J., and Samuel Epstein 1978 Influence of Diet on the Distribution of Carbon Isotopes in Animals. Geochimica et Cosmochimica Acta 42:495–506. 1981 Influence of Diet on the Distribution of Nitrogen Isotopes in Animals. Geochimica et Cosmochimica Acta 45:341–351. Digangi, Elizabeth A., and Megan K. Moore (editors) 2013 Research Methods in Human Skeletal Biology. Academic Press, Amsterdam. DiLeonardis, Lisa, and George F. Lau 2004 Life, Death, and Ancestors. In Andean Archaeology, edited by Helanie Silverman, pp. 77–115. Blackwell, Malden, MA. Dillehay, Tom D. 2001 Town and Country in Late Moche Times: A View from Two Northern

3 9 8   Reference List Valleys. In Moche Art and Archaeology in Ancient Peru, edited by Joanne Pillsbury, pp. 259–284. National Gallery of Art, Washington, DC. Dillehay, Tom D. (editor) 2011 From Foraging to Farming in the Andes: New Perspectives on Food Production and Social Organization. Cambridge University Press, Cambridge, UK. Dillehay, Tom D., and Lautaro Núñez 1988 Camelids, Caravans, and Complex Societies in the South-Central Andes. In Recent Studies in Pre Columbian Archaeology, edited by Nicholas J. Saunders and Oliver de Montmollin, pp. 603–633. BAR International Series 421, Oxford, UK. Dillon, R. 2013 Skeletal Remains from Huaca Santa Clara and Huaca Gallinazo. Manuscript on file with Jean-François Millaire, director, Virú Polity Project, London, Ontario. Dodson, Peter 1980 The Progress of Taphonomy. Science 210:631–632. Donley, Colleen Marie 2008 Late Moche Pit Burials from San José de Moro in Social and Political Perspective. In Arqueologia Mochica: Nuevos Enfoques, edited by Luis Jaime Castillo, Helaine Bernier, Julio Rucabado, and Gregory Lockard, pp. 119–130. Fondo Editorial de la Pontificia Universidad Católica del Perú, Lima. Donnan, Christopher B. 1973 Moche Occupation of the Santa Valley, Peru. University of California Press, Los Angeles. 1976 Moche Art and Iconography. Museum of Cultural History, University of California, Los Angeles. 1978 Moche Art of Peru: Pre-Columbian Symbolic Communication. Museum of Cultural History, University of California, Los Angeles. 1982 La Caza del Venado en el Arte Mochica. Revista del Museo Nacional 46:235–249. 1986a The Huaca 1 Complex. In The Pacatnamú Papers, Volume 1, edited by Christopher B. Donnan and Guillermo Cock, pp. 63–84. Fowler Museum of Cultural History, University of California, Los Angeles. 1986b Introduction. In The Pacatnamú Papers, Volume 1, edited by Christopher B. Donnan and Guillermo Cock, pp. 19–26. Fowler Museum of Cultural History, University of California, Los Angeles. 1988 Iconography of the Moche: Unraveling the Mystery of the Warrior-Priest. In National Geographic 174:551–555. 1989 En Busca de Naylamp: Chotuna, Chornancap, y el Valle de Lambayeque. In Lambayeque, edited by José Antonio de Lavalle, pp. 105–136. Coleccion Arte y Tesoros del Peru. Banco del Credito, Lima. 1990a An Assessment of the Validity of the Naymlap Dynasty. In The Northern Dynasties: Kingship and Statecraft in Chimor, edited by Michael E. Moseley and Alana Cordy-Collins, pp. 243–274. Dumbarton Oaks, Washington, DC.

Reference List  3 9 9 1990b The Chotuna Friezes and the Chotuna-Dragon Connection. In The Northern Dynasties: Kingship and Statecraft in Chimor, edited by Michael Moseley and Alana Cordy-Collins, pp. 234–274. Dumbarton Oaks, Washington, DC. 1992 Ceramics of Ancient Peru. Fowler Museum of Cultural History, Los Angeles. 1994 The Tombs of Ancient Peruvian Rulers: Myth Becomes Reality. Americas 43:14–22. 1995 Moche Funerary Practice. In Tombs for the Living: Andean Mortuary Practices, edited by Tom D. Dillehay, pp. 111–160. Dumbarton Oaks, Washington, DC. 1997a Deer Hunting and Combat: Parallel Activities in the Moche World. In The Spirit of Ancient Peru: Treasures from the Museo Arqueológico Rafael Larco Herrera, edited by Kathleen Berrin, pp. 51–59. Thames and Hudson, London. 1997b Introduction. In The Pacatnamu Papers, Volume II: The Moche Occupation, edited by Christopher Donnan and Guillermo Cock, pp. 9–16. Fowler Museum of Cultural History, Los Angeles. 2001 Moche Burials Uncovered. National Geographic 199:58–73. 2003 Tumbas con Entierros en Miniatura: Un Nuevo Tipo Funerario Moche. In Moche Hacia el Final Del Milenio, Tomo 1, edited by Santiago Uceda and Elías Mujica, pp. 43–78. Universidad Nacional de Trujillo y Pontificia Universidad Católica del Perú, Lima. 2004 Moche Portraits from Ancient Peru. University of Texas Press, Austin. 2007 Moche Tombs at Dos Cabezas. Cotsen Institute of Archaeology Press, Los Angeles. 2009 The Gallinazo Illusion. In Gallinazo: An Early Cultural Tradition on the Peruvian North Coast, edited by Jean-François Millaire and Magali Morlion, pp. 17–32. Cotsen Institute of Archaeology Press, Los Angeles. 2010 Moche State Religion. In New Perspectives on Moche Political Organization, edited by Jeffrey Quilter and Luis Jaime Castillo, pp. 47–69. Dumbarton Oaks, Washington, DC. 2011 Moche Substyles: Keys to Understanding Moche Political Organization. Boletín del Museo Chileno de Arte Precolombino 16:105–118. 2012 Chotuna and Chornancap: Excavating an Ancient Peruvian Legend. UCLA Cotsen Institute of Archaeology Press, Los Angeles. Donnan, Christopher B., and Luis Jaime Castillo 1992 Finding the Tomb of a Moche Priestess. Archaeology 45(6):38–42. 1994 Excavaciones de Tumbas de Sacerdotisas Moche en San José de Moro, Jequetepeque. In Moche, Propuestas y Perspectivas. Actas del Primer Coloquio sobre la Cultura Moche, edited by Santiago Uceda and Elias Mujica, pp. 415–424. Universidad Nacional de La Libertad, Trujillo, Peru. Donnan, Christopher B., and Guillermo A. Cock (editors) 1997 The Pacatnamu Papers, Volume 2: The Moche Occupation. Fowler Museum of Cultural History, University of California, Los Angeles. Donnan, Christopher B., and Leonard J. Foote 1978 Appendix 2. Child and Llama burials from Huanchaco. In Ancient Burial

4 00  Reference List Patterns of the Moche Valley, Peru, edited by Christopher B. Donnan and Carol J. Mackey, pp. 399–408. University of Texas Press, Austin. Donnan, Christopher B., and Carol J. Mackey 1978 Ancient Burial Patterns of the Moche Valley, Peru. University of Texas Press, Austin. Donnan, Christopher B., and Donna McClelland 1979 The Burial Theme in Moche Iconography. Studies in Pre-Columbian Art and Archaeology 21. Dumbarton Oaks, Washington, DC. 1999 Moche Fineline Painting: Its Evolution and Its Artists. Fowler Museum of Cultural History, University of California, Los Angeles. Dransart, Penelope Z. 2002 Earth, Water, Fleece, and Fabric: An Ethnography and Archaeology of Andean Camelid Herding. Routledge, London. Drusini, Andrea G., and Jose Pablo Baraybar 1991 Anthropological Study of Nasca Trophy Heads. Homo 41:251–265. Duday, Henri 1997 Antropología Biológica “De Campo,” Tafonomía y Arqueología de la Muerte. In El Cuerpo Humano y Su Tratamiento Mortuorio, edited by Elsa Malvido, Grégory Pereira, and Vera Tiesler, pp. 91–126. INAH-CEMCA, Mexico City. 2006 L’archéothantologie ou l’archéologie de la Mort (Archaeothanatology or the Archaeology of Death). Translation by Christopher Knüsel. In Social Archaeology of Funerary Remains, edited by Rebecca Gowland and Christopher Knüsel, pp. 30–56. Oxbow, Oxford, UK. 2009 The Archaeology of the Dead: Lectures in Archaeoathantology. Translation by Ana Marie Cipriani and John Pearch. Oxbow, Oxford, UK. Dufour, Elise, Nicolas Goepfert, Belkys Gutiérrez Léon, Claude Chauchat, Régulo Franco Jordan, and Segundo Vásquez Sánchez 2014 Pastoralism in Northern Peru During Pre-Hispanic Times: Insights from the Mochica Period (100–800 AD) Based on Stable Isotopic Analysis of Domestic Camelids. PLoS One 9:e87559. Duncan, William N. 2005 Understanding Veneration and Violation in the Archaeological Record. In Interacting with the Dead: Perspectives on Mortuary Archaeology for the New Millennium, edited by Gordon F. M. Rakita, Jane E. Buikstra, Lane Anderson Beck, and Sloan R. Williams, pp. 207–227. University Press of Florida, Gainesville. Duncan, William, and Kevin Schwartz 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 Ann Osterholtz, Kathryn Baustian, and Debra Martin, pp. 149–170. Springer, New York. Duviols, Pierre 1976 La Capacocha. Allpanchis: Revista del Instituto Pastoral Andino 9:11–57. Eeckhout, Peter, and Lawrence S. Owens 2008 Human Sacrifice at Pachacamac. Latin American Antiquity 19:375–398.

Reference List  4 01 Efremov, Ivan 1940 Taphonomy: A New Branch of Paleontology. Pan-American Geologist 74:81–93. Elera, Carlos G. 1986 Investigaciones Sobre Patrones Funerarios en el Sitio Formativo del Morro de Eten, Valle de Lambayeque, Costa Norte del Peru. Unpublished BA thesis, Pontificia Universidad Católica del Perú, Lima. 1998 The Puémape Site and the Cupisnique Culture: A Case Study on the Origins and Development of Complex Society in the Central Andes, Peru. Unpublished PhD dissertation, Department of Archaeology, University of Calgary, Canada. Evans-Pritchard, Edward 1940 The Nuer: A Description of the Modes of Livelihood and Political Institutions of a Nilotic People. Clarendon Press, Oxford, UK. 1956 Nuer Religion. Clarendon Press, Oxford, UK. Eveleth, Phyllis B, and James M. Tanner 1990 World-Wide Variation in Human Growth. 2nd ed. Cambridge University Press, Cambridge, UK. Faherty, Robert L. 2011 Sacrifice. In Encyclopædia Britannica online. Available at: http://www.britan nica.com/EBchecked/topic/515665/sacrifice. Accessed January 10, 2013. Farmer, Paul 2003 Pathologies of Power: Health, Human Rights, and the New War on the Poor. University of California Press, Berkeley. Farnum, Julie 2002 Biological Consequences of Social Inequalities in Prehistoric Peru. Unpublished PhD dissertation, Department of Anthropology, University of Missouri, Columbia. Faulkner, David K. 1986 The Mass Burial: An Entomological Perspective. In The Pacatnamu Papers, Volume 1, edited by Christopher B. Donnan and Guillermo A. Cock, pp. 145–150. Fowler Museum of Cultural History, Los Angeles. Feldman, J. A. 1977 The Architecture of Nias, Indonesia, with Special Reference to Bowomataluo Village. Unpublished PhD dissertation, Faculty of Philosophy, Columbia University, New York. Fernández, Jorge, Héctor O. Panarello, and Juan Schobinger 1999 The Inka Mummy from Mount Aconcagua: Decoding the Geographic Origin of the “Messenger to the Deities” by Means of Stable Carbon, Nitrogen, and Sulfur Isotope Analysis. Geoarchaeology 14(1):27–46. Fernández, Julio César 2004 Sinto: Señorío e Identidad en la Costa Norte Lambayecana. Chiclayo: COPROTUR. Fernández, Marco Antonio 2011 Informe Final—Temporada 2010: Proyecto de Investigación Arqueológica “La Pava.” Technical report on file at the Museo Nacional

4 02   Reference List de Arqueología y Etnografía Hans Henrich Brüning, Lambayeque, Perú. Fernández-Jalvo, Yolanda J., Carlos Díez, Isabel Cáceres, and Jordi Rosell 1999 Human Cannibalism in the Early Pleistocene of Europe (Gran Dolina, Sierra de Atapuerca, Burgos, Spain). Journal of Human Evolution 37:591–622. Fieller, Nick, and A. Turner 1982 Number Estimation in Vertebrate Samples. Journal of Archaeological Science 9:49–62. Finucane, Brian 2007 Mummies, Maize, and Manure: Multi-Tissue Stable Isotope Analysis of Late Prehistoric Human Remains from the Ayacucho Valley, Peru. Journal of Archaeological Science 34:2115–2124. Finucane, Brian, Patricia Maita Agurto, and William H. Isbell 2006 Human and Animal Diet at Conchopata, Peru: Stable Isotope Evidence for Maize Agriculture and Animal Management Practices during the Middle Horizon. Journal of Archaeological Science 33:1766–1776. Firth, Raymond 1963a Offering and Sacrifice: Problems of Organization. The Journal of the Royal Anthropological Institute of Great Britain and Ireland 93:12–24. 1963b We the Tikopia: A Sociological Study of Kinship in Primitive Polynesia. Beacon Press, Boston. Flannery, Kent V., Joyce Marcus, and Robert G. Reynolds 1989 The Flocks of Wamani: A Study of the Llama Herders on the Punas of Ayacucho, Peru. Academic Press, San Diego. Flores-Ochoa, Jorge A. 1979 Pastoralists of the Andes: The Alpaca Herders of Paratía. Translation by R. Bolton. Institute for the Study of Human Issues, Philadelphia. Fogel, Heidy P. 1993 Settlements in Time: A Study of Social and Political Development during the Gallinazo Occupation of the North Coast of Peru. Unpublished PhD dissertation, Department of Anthropology, Yale University, New Haven. Fogel, Marilyn L., Noreen Tuross, and Douglas W. Owsley 1989 Nitrogen Isotope Tracers of Human Lactation in Modern and Archaeological Population. Carnegie Institution Washington Yearbook 88:111–117. Ford, James A. 1949 Cultural Dating of Prehistoric Sites in Viru Valley, Peru. Anthropological Papers of the American Museum of Natural History 43:31–78. Fortes, Gloria G., Camilla F. Speller, Michael Hofreiter, and Turi E. King 2013 Phenotypes from Ancient DNA: Approaches, Insights and Prospects. BioEssays 35(8):690–695. Fortes, Meyer 1980 Preface. In Sacrifice, edited by M.  F.  C. Bourdillon and Meyer Fortes, pp. v–xix. Academic Press, New York.

Reference List  4 03 Fowler, Murray E. 1998 Medicine and Surgery of South American Camelids. Iowa State University Press, Ames. Fowler, William R. 1984 Late Preclassic Mortuary Patterns and Evidence for Human Sacrifice at Chalchuapa, El Salvador. American Antiquity 49:603–618. Frame, Mary 2001 Blood, Fertility, and Transformation: Interwoven Themes in the Paracas Necropolis Embroideries. In Ritual Sacrifice in Ancient Peru, edited by Elizabeth P. Benson and Anita G. Cook, pp. 55–92. University of Texas Press, Austin. Franco, Régulo 2008 La Señora de Cao. In Señores de los Reinos de la Luna, edited by Krzysztof Makowski, pp. 280–287. Banco de Crédito, Lima. Franco, Régulo, and César Gálvez 2005 Muerte, Identidades, y Prácticas Funerarias Post-Mochicas en El Complejo El Brujo, Valle de Chicama, Costa Norte del Peru. Corriente Arqueológica 1:79–118. Franco, Régulo J., César Galvez M., and Segundo Vásquez S. 1994 Proyecto Arqueológico Complejo El Brujo, Informe Final, Temporada 1994. Convenio Fundación Augusto N. Wiese, Instituto Regional de Cultura— La Libertad, Universidad Nacional de Trujillo. 1998 Desentierro Ritual de una Tumba Moche: Huaca Cao Viejo. Revista Arqueológica Sian 3:9–18. 2001 La Huaca Cao Viejo en el Complejo El Brujo: Una Contribución al Estudio de los Mochicas en el Valle de Chicama. Arqueologicas 25:123–173. Frayer, David W. 1997 Ofnet: Evidence for a Mesolithic Massacre. In Troubled Times: Violence and Warfare in the Past, edited by Debra L. Martin and David W. Frayer, pp. 181–216. Gordon and Breach, Amsterdam. Fry, Douglas P. 2005 The Human Potential for Peace: An Anthropological Challenge to Assumptions about War and Violence. Oxford University Press, Oxford, UK. Fry, Douglas P. (editor) 201 War, Peace, and Human Nature: The Convergence of Evolutionary and Cultural Views. Oxford University Press, Oxford, UK. Fuchs, Peter R. 1997 Nuevos Datos Arqueométricos para la Historia de Ocupación de Cerro Sechín: Período Lítico al Formativo. In Arqueologia Peruana 2: Arquitectura y Civilización en los Andes Prehispánicos, edited by Elisabeth Bonnier and Henning Bischoff, pp. 145–161. Sociedad Arqueológica PeruanoAlemana, Reiss-Museum, Manheim. Funabiki, Takeo 1981 On Pigs of the Mbotgote in Malekula. In Vanuatu: Politics, Economics, and Ritual in Island Melanesia, edited by Michael R. Allen, pp. 173–188. Academic Press, New York.

4 04   Reference List Gagné, Gérard 2009 Gallinazo Disposal of the Dead and Manipulation of Human Remains at El Castillo de Santa. In Gallinazo: An Early Cultural Tradition on the Peruvian North Coast, edited by Jean-François Millaire and Magali Morlion, pp. 207–222. Cotsen Institute of Archaeology Press, Los Angeles. Gaither, Catherine M. 2004 Estudio del Crecimiento y Desarrollo Dental Humano en la Prehistoria de la Costa de Perú: Implicaciones Paleopatológicas. Archaeobios 1:10–14. 2008 A Growth and Development Study of Coastal Prehistoric Peruvian Populations. Unpublished PhD dissertation, Department of Anthropology, Tulane University, New Orleans. Gaither, Catherine M., and Mellisa S. Murphy 2012 Consequences of Conquest? The Analysis and Interpretation of Subadult Trauma at Puruchuco-Huaquerones, Peru. Journal of Archaeological Science 39:467–478. Gaither, Catherine M., Jonathan Bethard, Jonathan D. Kent, Victor Vasquez, Teresa Rosales, and Richard Busch 2010 Strange Harvest: A Discussion of Sacrifice and Missing Body Parts on the North Coast of Peru. Andean Past 9:177–194. Gaither, Catherine M., Jonathan D. Kent, Víctor Vásquez Sánchez, and Teresa Rosales Tam 2008 Mortuary Practices and Human Sacrifice in the Middle Chao Valley of Peru: Their Interpretation in the Context of Andean Mortuary Patterning. Latin American Antiquity 19:107–121. Galloway, Allison 1997 Process of Decomposition: A Model from the Arizona-Sonoran Desert. In Forensic Taphonomy: The Postmortem Fate of Human Remains, edited by William D. Haglund and Marcella H. Sorg, pp. 139–150. CRC Press, Boca Raton, Florida. 1999 Fracture Patterns and Skeletal Morphology: Introduction and the Skull. In Broken Bones: Anthropology Analysis of Blunt Force Trauma, edited by Galloway, Allison, p. 63. Charles C. Thomas, Springfield, Illinois. Galtung, Johan 1969 Violence, Peace, and Peace Research. Journal of Peace Research 6:167–191. Gálvez Mora, César, and Jesus Briceño Rosario 2001 The Moche in the Chicama Valley. In Moche Art and Archaeology in Ancient Peru, edited by Joanne Pillsbury, pp. 141–157. National Gallery of Art, Washington, DC. Garcilaso de la Vega, “El Inca” 1945 [1609] Comentarios Reales de los Incas. Editoral Rosenblatt, Buenos Aires. Gell, Alfred
 1998 Art and Agency: An Anthropological Theory. Clarendon, Oxford, UK. Genovés, Santiago 1967 Proportionality of Long Bones and Their Relation to Stature among Mesoamericans. American Journal of Physical Anthropology 26:67–78.

Reference List  4 05 Gentile, Margarita 1996 Dimensión Sociopolitíca y Religiosa de la Capacocha del Cerro Aconcagua. Bulletin del Institute Frances des Etudes Andines 25:43–90. Gibson, Thomas 1988 Meat Sharing as a Political Ritual: Forms of Transaction Versus Modes of Subsistence. In Hunters and Gatherers 2: Property, Power, and Ideology, edited by Tim Ingold, David Riches, and James Woodburn, pp. 165–179. Berg, New York. Gilby, A. R., and J. W. McKellar 1970 The Composition of the Empty Puparia of a Blowfly. Journal of Insect Physiology 16:1517–1529. Girard, René 1977 [1972] Violence and the Sacred [La Violence et le Sacre]. Johns Hopkins University Press, Baltimore. Glassman, David M., and Suzanna E. Dana 1992 Handedness and the Bilateral Asymmetry of the Jugular Foramen. Journal of Forensic Sciences 37:140–146. Goepfert, Nicolas 2010 The Llama and the Deer: Dietary and Symbolic Dualism in the Central Andes. Anthropozoologica 45:25–45. 2012 New Zooarchaeological and Funerary Perspectives on Mochica culture (A.D. 100–800), Peru. Journal of Field Archaeology 37:104–120. Goepfert, Nicolas, Elise Dufour, Belkys Gutiérrez, and Claude Chauchat 2013 Origen Geográfico de Camélidos en el Periodo Mochica (100–800 AD) y Análisis Isotópico Secuencial del Esmalte Dentario: Enfoque Metodológico y Aportes Preliminares. Bulletin de l’Institut Français d’Études Andines 42:1–24. González, Yolotl 1985 El Sacrificio Humano Entre Los Mexicas. Fondo de Cultura Económica, Mexico City. Goodman, Alan H., and George J. Armelagos 1985 Factors Affecting the Distribution of Enamel Hypoplasias within the Human Permanent Dentition. American Journal of Physical Anthropology 68:479–493. Goodman, Alan H., George J. Armelagos, and Jerome C. Rose 1980 Enamel Hypoplasia as Indicators of Stress in Three Prehistoric Populations from Illinois. Human Biology 52:515–528. Goodman, Alan H., and Debra L. Martin 2002 Reconstructing Health Profiles from Skeletal Remains. In Backbone of History: Health and Nutrition in the Western Hemisphere, edited by Richard H. Steckel and Jerome C. Rose, pp. 11–60. Cambridge University Press, Cambridge, UK. Goodman, Alan H., and Jerome C. Rose 1991 Dental Enamel Hypoplasias as Indicators of Nutritional Status. In Advances in Dental Anthropology, edited by Marc A. Kelley and Clark Spencer Larsen, pp. 279–293. Wiley-Liss, New York.

4 06   Reference List Gose, Peter 1994 Deathly Waters and Hungry Mountains: Agrarian Ritual and Class Formation in an Andean Town. University of Toronto Press, Toronto. Gowland Rebecca, and Christopher Knüsel 2006 Introduction. In Social Archaeology of Funerary Remains, edited by Rebecca Gowland and Christopher Knüsel, pp. ix–xiv. Oxbow, Oxford, UK. Gravlee, Clarence 2009 How Race Becomes Biology: Embodiment of Social Inequality. American Journal of Physical Anthropology 139:47–57. Grayson, Donald 1978 Minimum Numbers and Sample Size in Vertebrate Faunal Analysis. American Antiquity 43:53–65. Green, Marc A. 1973 Morbid Anatomical Findings in Strangulation. Forensic Science 2:317–323. Green, Miranda A. 1992 Animals in Celtic Life and Myth. Routledge, London. 2001 Dying for the Gods: Human Sacrifice in Iron Age and Roman Europe. Tempus, Charleston, South Carolina. Gumerman, George, IV 1994 Corn for the Dead: The Significance of Zea mays in Moche Burial Offerings. In Corn and Culture in the Prehistory New World, edited by Sissel Johannessen and Christine A. Hastorf, pp. 399–410. Westview, Boulder, Colorado. Haas, Jonathan 1985 Excavations on Huaca Grande: An Initial View of the elite at Pampa Grande. Journal of Field Archaeology 12:391–409. Haglund, William H. 1992 Contribution of Rodents to Postmortem Artifacts of Bone and Soft Tissues. Journal of Forensic Sciences 37:1459–1465. 1997 Rodents and Human Remains. In Forensic Taphonomy: The Postmortem Fate of Human Remains, edited by William D. Haglund and Marcella H. Sorg, pp. 405–414. CRC Press, Boca Raton, Florida. Haglund, William H., Donald T. Reay, and Daris R. Swindler 1988 Tooth Mark Artifacts and Survival of Bones in Animal Scavenged Human Remains. Journal of Forensic Sciences 33:985–997. 1989 Canid Scavenging/Disarticulation Sequence of Human Remains in the Pacific Northwest. Journal of Forensic Sciences 34:587–606. Haglund, William, and Marcella H. Sorg 1997 Introduction to Forensic Taphonomy. In Forensic Taphonomy: The Postmortem Fate of Human Remains, edited by William D. Haglund and Marcella H. Sorg, pp. 1–9. CRC Press, Boca Raton, Florida. Haglund, William, and Marcella Sorg (editors) 2002 Advances in Forensic Taphonomy: Method, Theory and Archaeological Perspectives. CRC Press, Boca Raton, Florida. Hamayon, Roberte N. 1990 La Chasse à l’Áme, Esquisse d’une Théorie due Chamanisme Sibérien. Société d’Ethnologie, Nanterre, France.

Reference List  4 07 Hamilakis, Yannis, and Eleni Konsolaki 2004 Pigs for the Gods: Burnt Animal Sacrifices as Embodied Rituals at a Mycenaean Sanctuary. Oxford Journal of Archaeology 23:135–151. Handley, Linda L., A. T. Austin, G. R. Stewart, D. Robinson, C. M. Scrimgeour, J. A. Raven, T. H. E. Heaton, and S. Schmidt 1999 The 15N Natural Abundance (δ15N) of Ecosystem Samples Reflects Measures of Water Availability. Australian Journal of Plant Physiology 26:185–199. Harman, Mary, Theya I. Molleson, and J. L. Price 1981 Burials, Bodies and Beheadings in Romano-British and Anglo-Saxon Cemeteries. Bulletin of the British Museum (Natural History) 35:145–188. Hasu, Päivi 2009 For Ancestors and God: Rituals of Sacrifice among the Chagga of Tanzania. Ethnology 48:195–213. Hayashida, Francis 1998 New Insights into Inka Pottery Production. In Andean Ceramics: Technology, Organization, and Approaches, edited by Izumi Shimada, pp. 313–335. Museum Applied Science Center for Archaeology and University of Pennsylvania Museum of Archaeology and Anthropology, Philadelphia. 2006 The Pampa de Chaparrí: Water, Land, and Politics on the North Coast of Peru. Latin American Antiquity 17:243–263. Haynes, Gary 1980 Evidence of Carnivore Gnawing on Pleistocene and Recent Mammalian Bones. Paleobiology 6:341–351. 1983 A Guide for Differentiating Mammalian Carnivore Taxa Responsible for Gnaw Damage to Herbivore Limb Bones. Paleobiology 9:164–172. Hecker, Gisella, and Wolfgang Hecker 1992 Ofrendas de Huesos Humanos y Uso Repetido de Vasijas en el Culto Funerario de la costa Norperuana. Gaceta Aqueológica Andina 6:33–53. Hedges, Robert E.  M., John G. Clement, David L. Thomas, and Tamsin C. O’Connell 2007 Collagen Turnover in the Adult Femoral Mid-Shaft: Modeled from Anthropogenic Radiocarbon Tracer Measurements. American Journal of Physical Anthropology 133:808–816. Heesterman, J. C. 1993 The Broken World of Sacrifice: An Essay in Ancient Indian Ritual. University of Chicago Press, Chicago. Helfrich, Heinrich 1973 Menschenopfer und Toetungsrituale im Kult der Maya, Berlin, Mann Verlag. Hesse, Brian, Paula Wapnish, and Jonathan Greer 2012 Scripts of Animal Sacrifice in Levantine Culture-History. In Sacred Killing: The Archaeology of the Ancient Near East, edited by Anne Porter and Glenn M. Schwartz, pp. 217–236. Eisenbrauns, Winona Lake, Indiana. Hewitt, Barbara R. 2013 Foreigners among the Dead at Túcume, Peru: Assessing Residential Mobility Using Isotopic Tracers. Unpublished PhD dissertation, The University of Western Ontario, London, Ontario.

4 08   Reference List Hewitt, Barbara R., Christine D. White, J. Marla Toyne, Fred J. Longstaffe, and B. J. Fryer 2008 The Aqlla of Tucume? Biogeochemical and Bioarchaeological Analyses of 19 Individuals Buried at Huaca Larga. Paper presented at the 73rd Annual Meeting of the Society for American Archaeology, Vancouver, British Columbia, March 26–30. Heyerdahl, Thor, Daniel H. Sandweiss, and Alfredo Narvaéz 1995 Pyramids of Túcume: The Quest for Peru’s Forgotten City. Thames and Hudson, London. Hill, Andrew P. 1979 Disarticulation and Scattering of Mammal Skeletons. Paleobiology 5:261–274. Hill, Erica 1998 Death as a Rite of Passage: The Iconography of Moche Burial Theme. Antiquity 72:528–538. 2003 Sacrificing Moche Bodies. Journal of Material Culture 8:285–299. 2005 The Body Intact, the Body in Pieces: Foucault and the Moche Prisoner. In Art for Archaeology’s Sake: Material Culture and Style Across the Disciplines, edited by Andrea Waters-Rist, Christine Cluney, Calla McNamee, and Larry Steinbrenner, pp. 124–131. University of Calgary, Calgary, Alabama. 2006 Moche Skulls in Cross-Cultural Perspective. In Skull Collection, Modification, and Decoration, edited by Michelle Bonogofsky, pp. 91–100. Archaeopress, Oxford, UK. Hillson, Simon 2008 The Current State of Dental Decay. Technique and Application in Dental Anthropology, edited by Joel D. Irish and Greg C. Nelson, pp. 111–135. Cambridge University Press, Cambridge, UK. Hocquenghem, Anne-Marie 1980 L’iconographie Mochica et les représentations de supplices. Journal de la Société des Américanistes 67:249–260. 1983 Les Cerfs et les Morts dans l’iconographie Mochica. Journal de la Société des Américanistes 69:71–84. 1987 Iconografía Mochica. Fondo Editorial de la Pontificia Universidad Católica del Perú, Lima. 2008 Sacrifices and Ceremonial Calendars in Societies of the Central Andes: A Reconsideration. In The Art and Archaeology of the Moche: An Ancient Andean Society of the Peruvian North Coast, edited by Steve Bourget and Kimberly L. Jones, pp. 23–42. University of Texas Press, Austin. Hoffmann, Beatrix 2007 Posibilidades y Limitaciones para la Reconstrucción y Recontextualización de la Colección Gretzer del Museo Etnológico de Berlin. Baessler-Archiv 55:165–178. Hoskins, Janet 1993 Violence, Sacrifice, and Divination: Giving and Taking Life in Eastern Indonesia. American Ethnologist 20:159–178.

Reference List  4 09 Houston, Stephen, David Stuart, and Karl A. Taube 2006 The Memory of Bones. Body, Being, and Experience Among the Classic Maya. University of Texas, Austin. Howells, W. W. 1989 Skull Shapes and the Map: Craniometric Analyses in the Dispersion of Modern Homo. Peabody Museum of Archaeology and Ethnology, Harvard University, Cambridge, Massachusetts. Hubbard, Amelia R., Debbie Guatelli-Steinberg, and Joel D. Irish 2015 Do Nuclear DNA and Dental Nonmetric Data Produce Similar Reconstructions of Population History? An Example from Modern Coastal Kenya. American Journal of Physical Anthropology 157(2):295–304. Hubert, Henri, and Marcel Mauss 1964 [1898] Sacrifice: Its Nature and Function. Translation by W. D. Halls. University of Chicago Press, Chicago. Huchet, Jean-Bernard, and Bernard Greenberg 2010 Flies, Mochicas, and Burial Practices: A Case Study from Huaca de la Luna, Peru. Journal of Archaeological Science 37:2846–2856. Huchet, Jean-Bernard, D. Deverly, Belkys Gutierrez, and Claude Chauchat 2011 Taphonomic Evidence of a Human Skeleton Gnawed by Termites in a Moche-Civilization Grave at Huaca de la Luna, Peru. International Journal of Osteoarchaeology 21:92–102. Hudson Museum 1998 Empires Emerging: Collecting the Peruvian Past: September 20, 1997–April 26, 1998. University of Maine, Orono. Hughes, Toby, and Grant C. Townsend 2013 Twin and Family Studies of Human Dental Crown Morphology: Genetic, Epigenetic, and Environmental Determinants of the Modern Human Dentition. In Anthropological Perspectives on Tooth Morphology: Genetics, Evolution, Variation, edited by G. Richard Scott and Joel D. Irish, pp. 31–68. Cambridge University Press, Cambridge, UK. Humphrey, Caroline, and James Laidlaw 2007 Sacrifice and Ritualization. In The Archaeology of Ritual, edited by Evangelos Kyriakidis, pp. 225–276. UCLA Cotsen Institute of Archaeology Press, Los Angeles. Hurtubise, Jenna, Derrick Nuesmeyer, Haagen Klaus, José Pinilla, Ana Alva, and Carlos Elera 2014 Trauma or Taphonomy? A Possible Case of Cranial Impalement in the Sacrificial Context of Matrix 101, Huaca Las Ventanas, Peru. Poster presented at the 40th Annual Meeting of the Paleopathology Association, Calgary, Alberta, Canada. IBM Corp. 2012 IBM SPSS Statistics for Windows, Version 21.0. IBM Corporation, Armonk, New York. Imbelloni, Juan 1933 Los Pueblos Deformadores de los Andes. La Deformación Intencional

4 1 0  Reference List de la Cabeza Como Arte y como Elemento Diagnóstico de las culturas. Anales del Museo Argentino de Ciencias Naturales 37:209–253. Ingold, Tim 1987 The Appropriation of Nature: Essays on Human Ecology and Social Relations. University of Iowa Press, Iowa City. 1994 From Trust to Domination: An Alternative History of Human-Animal Relations. In Animals and Human Society: Changing Perspectives, edited by Aubrey Manning and James Serpell, pp. 1–22. Routledge, New York. 2007 Materials Against Materiality. Archaeological Dialogues 14:1–16. Insoll, Timothy 2010 Talensi Animal Sacrifice and Its Archaeological Implications. World Archaeology 42:231–244. Isaakidou, Valasia, Paul Halstead, Jack Davis, and Sharon Stocker 2002 Burnt Animal Sacrifice at the Mycenaean “Palace of Nestor,” Pylos. Antiquity 76:86–92. Isbell, William H. 1987 Conchopata, Ideological Innovator in Middle Horizon. Ñawpa Pacha 22:91–126. Isbell, Willam H., and Anita G. Cook 2002 A New Perspective on Conchopata and the Andean Middle Horizon. In Andean Archaeology II: Art, Landscape, and Society, edited by Helaine Silverman and William H. Isbell, pp. 249–305. Kluwer Academic/Plenum, New York. Iserson, Kenneth V. 1984 Strangulation: A Review of Ligature, Manual, and Postural Neck Compression Injuries. Annals of Emergency Medicine 13:179–185. Jackson, Margaret A. 2004 The Chimu Sculptures of Huacas Tacaynamo and El Dragon, Moche Valley, Peru. Latin American Antiquity 15:298–322. 2008 Moche Art and Visual Culture in Ancient Peru. University of New Mexico Press, Albuquerque. Jamieson, James B. 1983 An Examination of Prisoner-Sacrifice and Cannibalism at the St. Lawrence Iroquoian Roebuck Site. Canadian Journal of Archaeology 7:159–175. Jean, Guilaine, and Jean Zammit 2005 The Origins of War: Violence in Prehistory. Translation by Melanie Hersey. Blackwell, Malden, Massachusetts. Jelínek, Jaroslav 1993 Dismembering, Filleting, and Evisceration of Human Bodies in a Bronze Age Site in Moravia, Czech Republic. Anthropologie (Brno) 30(1–3):99–114. Jennings, Justin 2009 Catastrophe, Revitalization, and Religious Change on the Prehispanic North Coast of Peru. Cambridge Archaeological Journal 18:177–194. Jing, Yuan, and Rowan Flad 2005 New Zooarchaeological Evidence for Changes in Shang Dynasty Animal Sacrifice. Journal of Anthropological Archaeology 24:252–270.

Reference List  4 1 1 Jolly, Margaret 1984 The Anatomy of Pig Love: Substance, Spirit, and Gender in South Pentecost, Vanuatu. Canberra Anthropology 7(1–2):78–108. Jones, Kimberly L. 2010 Cupisnique Culture: The Development of Ideology in the Ancient Andes. Unpublished PhD dissertation, Department of Art and Art History, University of Texas, Austin. Judd, Margaret A. 2002 Ancient Injury Recidivism: An Example from the Kerma Period of Ancient Nubia. International Journal of Osteoarchaeology 12:89–106. 2004 Trauma in the City of Kerma: Ancient Versus Modern Injury Patterns. International Journal of Osteoarchaeology 14:34–51. 2008 The Parry Problem. Journal of Archaeological Science 35:1658–1666. Kaulike, Peter 1991 Chavín Art and Iconography. In Ancient Art of the Andean World, edited by Shozo Masudo and Izumi Shimada, pp. 47–69. Iwanami Shoten, Tokyo. Kellner, Corina M., and Margaret J. Schoeninger 2008 Wari’s Imperial Influence on Local Nasca Diet: The Stable Isotope Evidence. Journal of Anthropological Archaeology 27:226–243. Kelly, John E., James A. Brown, and Lucretia S. Kelly 2008 The Context of Religion at Cahokia: The Mound 34 Case. In Religion in the Material World, edited by Lars Fogelin, pp. 297–318. Center for Archaeological Investigations Occasional Papers No. 36. Southern Illinois University Press, Carbondale. Kent, Jonathan D., and Makota Kowta 1994 The Cemetery at Tambo Viejo, Acarí Valley, Peru. Andean Past 4:106–140. Kent, Jonathan D., Teresa E. Rosales Tham, and Victor F. Vásquez Sánchez 2003 Informe Final: Proyecto Manejo Ecosustentable y Desarrollo Cultural del Complejo Arqueológico Santa Rita “B”: Temporada 2002. On file with the Ministerio de Cultura, Lima. Kent, Jonathan D., Teresa E. Rosales Tham, Victor F. Vásquez Sánchez, and Jonathan D. Bethard 2006 Informe Final: Proyecto Manejo Ecosustentable y Desarrollo Cultural del Complejo Arqueológico Santa Rita “B”: Temporada 2006. On file with the Ministerio de Cultura, Lima. Kertzer, David I. 1988 Ritual, Politics, and Power. Yale University Press, New Haven, Connecticut. 1991 The Role of Ritual in State-Formation. In Religious Regimes and State Formation Perspectives from European Ethnohistory, edited by Eric R. Wolf, pp. 85–103. SUNY Press, Albany. Kieser, J. A. 1990 Adult Human Odontometrics. Cambridge University Press, Cambridge, UK. Kimmerle, Erin H., and Jose Pablo Baraybar 2008 Skeletal Trauma: Identification of Injuries Resulting from Human Rights Abuse and Armed Conflict. CRC Press, Boca Raton, Florida.

4 1 2   Reference List Klaus, Haagen D. 2003a Life and Death at Huaca Sialupe: The Mortuary Archaeology of a Middle Sicán Community, North Coast of Peru. Unpublished MA thesis, Department of Anthropology, Southern Illinois University, Carbondale. 2003b Field notes, Proyecto Arqueólogico Sicán Temporada 2003. On file with the author. 2008a Out of Light Came Darkness: Bioarchaeology of Mortuary Ritual, Health, and Ethnogenesis in the Lambayeque Valley Complex, North Coast Peru, A.D. 900–1750. Unpublished PhD dissertation, Department of Anthropology, The Ohio State University, Columbus. 2008b Human Skeletal Remains of Huaca el Pueblo and the Lord of Úcupe: Biological Variation, Health, and Moche Lordship. Technical report prepared for Steve Bourget, Huaca el Pueblo Archaeological Project. On file with the author. 2009 Sicán Human Sacrifices: Forms, Methods, Contexts, and Significance (in Japanese). In Precursor of the Inka Empire: The Golden Capital of Sicán, edited by Izumi Shimada, Ken-ichi Shinoda, and Masahiro Ono, pp. 311– 319. Tokyo Broadcasting System, Tokyo. 2011 Exploring Social and Ideological Change through the Archaeothanatology of Mortuary Ritual in Colonial Peru. Paper at presented at Death, Decay, and Discovery: An Interdisciplinary Workshop on Taphonomic Approaches to the Understanding of Burial Practice, Joukowski Institute for Archaeology and the Ancient World, Brown University, Providence, Rhode Island. 2012 The Bioarchaeology of Structural Violence: A Theoretical Model and a Case Study. In The Bioarchaeology of Violence, edited by Debra L. Martin, Ryan P. Harrod, and Ventura R. Perez, pp. 29–62. University Press of Florida, Gainesville. 2013 Field notes, Bioarchaeological Analysis of the Human Remains from Dos Cabezas, prepared for Steve Bourget and Kimberly L. Jones. On file with the author. 2014a La Población Muchik de la Cultura Sicán Medio: Una Primera Aproximación a un Sustrato Cultural Prehispánico Tardío del Valle de Lambayeque. In Cultura Sicán: Esplendor Preinca de la Costa Norte, edited by Izumi Shimada, pp. 235–257. Fondo Editorial del Congreso del Perú, Lima. 2014b A History of Violence in the Lambayeque Valley: Conflict and Death from the Late Pre-Hispanic Apogee to European Colonization of Peru (A.D. 900–1750). In The Routledge Handbook of the Bioarchaeology of Human Conflict, edited by Christopher J. Knüsel and Martin J. Smith, pp. 389– 414. Routledge, New York. 2014c Frontiers in the Bioarchaeology of Stress and Disease: Cross-Disciplinary Perspectives from Pathophysiology, Human Biology, and Epidemiology. American Journal of Physical Anthropology 155:294–308. 2014d Informe Final: Análisis Bioarqueológico de los Restos Humanos de Chornancap Tumbas 4, 5, y 6. Technical report on file at the Museo Nacional de Arqueología y Etnografía Hans Henrich Brüning, Lambayeque, Peru.

Reference List  4 1 3 Klaus, Haagen D., and Carlos Wester 2005 Health, Burial, and Social Structure at Úcupe (Zaña River Valley): A New Perspective on the Chimú of Ancient Northern Coastal Peru. American Journal of Physical Anthropology Supplement 38:131. Klaus, Haagen D., Clark Spencer Larsen, and Manuel E. Tam 2009 Economic Intensification and Degenerative Joint Disease: Life and Labor on the Postcontact North Coast of Peru. American Journal of Physical Anthropology 139:204–221. Klaus, Haagen D., Izumi Shimada, Ken-ichi Shinoda, and Sara Muno. In press Middle Sicán Mortuary Archaeology, Skeletal Biology, and Genetic Structure in Ancient South America. In Bones of Complexity: Bioarchaeological Case Studies in Social Organization and Skeletal Biology, edited by Haagen D. Klaus, Amanda Harvey, and Mark N. Cohen. University Press of Florida, Gainesville. Klaus, Haagen D., Jorge Centurión, and Manuel Curo 2010 Bioarchaeology of Human Sacrifice: Violence, Identity, and Ritual Killing at Cerro Cerrillos, Peru. Antiquity 84:1102–1122. Klaus, Haagen D., Kevin Reed, Jenna Hurtubise, Steven Nau, Angelina DeMarco, Alexis Meeks, José Pinilla, Ana Alva, and Carlos Elera 2013 The Bioarchaeological Enigmas of Matrix 101: A Mass Human Sacrifice of the Middle Sicán Period. Paper presented at the 40th Annual Meeting of the Paleopathology Association, Knoxville, Tennessee. Klaus, Haagen D., and Manuel E. Tam 2009a Surviving Contact: Biological Transformation, Burial, and Ethnogenesis in the Colonial Lambayeque Valley, Perú. In Bioarchaeology and Identity in the Americas, edited by Kelly J. Knudson and Christopher M. Stojanowski, pp. 124–152. University Press of Florida, Gainesville. 2009b Contact in the Andes: Bioarchaeology of Systemic Stress in Colonial Mórrope. American Journal of Physical Anthropology 138:356–368. 2010 Oral Health and the Postcontact Adaptive Transition: A Contextual Reconstruction of Diet in Mórrope, Peru. American Journal of Physical Anthropology 141:594–609. 2015 Requiem Aeternam? Archaeothanatology of Mortuary Ritual in Colonial Mórrope, North Coast of Peru. In Living with the Dead in the Andes, edited by Izumi Shimada and James Fitzsimmons, pp. 267–303. University of Arizona Press, Tucson. Knapp, Michael, and Michael Hofreiter 2010 Next Generation Sequencing of Ancient DNA: Requirements, Strategies, and Perspectives. Genes 1:227–243. Knight, John 2005 Introduction. In Animals in Person: Cultural Perspectives on HumanAnimal Intimacy, edited by John Knight, pp. 1–13. Berg, Oxford, UK. 2012 The Anonymity of the Hunt: A Critique of Hunting as Sharing. Current Anthropology 53:334–355. Knudson, Kelly J., Arthur E. Aufderheide, and Jane E. Buikstra 2007 Seasonality and Paleodiet in the Chiribaya Polity of Southern Peru. Journal of Archaeological Science 34:451–462.

4 14   Reference List Knudson, Kelly J., Kristin R. Gardella, and Jason Yaeger 2012 Provisioning Inka Feasts at Tiwanaku, Bolivia: The Geographic Origins of Camelids in the Pumapunku Complex. Journal of Archaeological Science 39:479–491. Knudson, Kelly J., and Christopher M. Stojanowski (editors) 2009 Bioarchaeology and Identity in the Americas. University Press of Florida, Gainesville. Knüsel, Christopher J. 2005 The Physical Evidence of Warfare-Subtle Stigmata? In Warfare, Violence, and Slavery in Prehistory, edited by Mike Parker Pearson and I.  N.  J. Thorpe, pp. 49–65. BAR International Series 1374. Archaeopress, Oxford, UK. 2010 Bioarchaeology: A Synthetic Approach/Bio-archéologie: Une Approche Synthétique. Bulletins de la Société d’Anthropologie de Paris (2010) 22:62–73. Knüsel, Christopher J., and Martin J. Smith (editors) 2014 The Routledge Handbook of the Bioarchaeology of Human Conflict. Routledge, New York. Kohn, Matthew J. 2010 Carbon Isotope Compositions of Terrestrial C3 Plants as Indicators of (Paleo)ecology and (Paleo)climate. Proceedings of the National Academy of Sciences 107:19691–19695. Kolata, Alan L. 1990 The Urban Concept of Chan Chan. In The Northern Dynasties: Kingship and Statecraft in Chimor, edited by Michael Moseley and Alana CordyCollins, pp. 107–144. Dumbarton Oaks, Washington, DC. 1996 Valley of the Spirits: A Journey into the Lost Realm of the Aymara. John Wiley and Sons, New York. 2006 Before and After Collapse: Reflections on the Regeneration of Social Complexity. In After Collapse: The Regeneration of Complex Societies, edited by Glenn M. Schwartz and John J. Nichols, pp. 208–221. University of Arizona Press, Tucson. Koschmieder, Klaus, and Catherine Gaither 2010 Tumbas de Guerreros Chachapoya en Abrigos Rocosos de la Provincia de Luya, Departamento de Amazonas. Arqueología y Sociedad 22:1–30. Kosok, Paul 1965 Life, Land, and Water in Ancient Peru. Long Island University Press, New York. Krieger, Nancy 2004 Embodying Inequality: Epidemiological Perspectives. Baywood, Amityville, New York. Kroeber, Alfred 1925 The Uhle Pottery Collections from Moche. University of California Publications in American Archaeology and Ethnology 21:191–234. Küchler, Susanne 2002 Malanggan: Art, Memory, and Sacrifice. Berg, Oxford, UK. Kuckelman, Kristin A., Ricky R. Lightfoot, and Debra L. Martin

Reference List  4 1 5 2002 The Bioarchaeology and Taphonomy of Violence at Castle Rock and Sand Canyon Pueblos, Southwestern Colorado. American Antiquity 67:486–513. Kunen, Julie L., Mary Jo Galindo, and Erin Chase 2002 Pits and Bones: Identifying Maya Ritual Behavior in the Archaeological Record. Ancient Mesoamerica 13:197–211. Kurin, Daneille 2014 Cranial Trauma and Cranial Modification in Post-Imperial Andahuaylas, Peru. In Bioarchaeological and Forensic Perspectives on Violence: How Death Is Interpreted from Skeletal Remains, edited by Debra L. Martin and Cheryl P. Anderson, pp. 239–260. Labarca, R., and F. Gallardo 2012 The Domestic Camelids (Cetartiodactyla: Camelidae) from the Middle Formative Cemetery of Topater 1 (Atacama Desert, Northern Chile): Osteometric and Palaeopathological Evidence of Cargo Animals. International Journal of Osteoarchaeology. 25(1):61–73. Lahren, Craig H., and Hugh E. Berryman 1984 Fracture Patterns and Status at Chucalissa (40SY1): A Biocultural Approach. Tennessee Anthropologist 9:15–21. Lambert, Patricia M., Brian R. Billman, and Banks L. Leonard 2000 Explanation Variability in Mutilated Human Bone Assemblages from the American Southwest: A Case Study from the Southern Piedmont of Sleeping Ute Mountain, Colorado. International Journal of Osteoarchaeology 10:49–64. Lambert, Patricia, Celeste Marie Gagnon, Brian R. Billman, M. Anne Katzenberg, Jose Carcelen, and Robert H. Tykot 2012 Bone Chemistry at Cerro Oreja: A Stable Isotope Perspective on the Development of a Regional Economy in the Moche Valley, Peru, During the Early Intermediate Period. Latin American Antiquity 23:144–166. Lapiner, Alan C. 1976 Pre-Columbian Art of South America. H. N. Abrams, New York. Larco Hoyle, R. 1941 Los Cupisniques. Casa Editora “La Crónica” y “Variedades,” S.A. ltda., Lima. 1945 La Cultura Virú. Sociedad Geográfica Americana, Buenos Aires. 1962 La Cultura Santa. Lit. Velarde S.A., Lima. 2001a Los Mochicas. Tomo I. Museo Arqueológico Rafael Larco Herrera, Lima. 2001b Los Mochicas. Tomo II. Museo Arqueológico Rafael Larco Herrera, Lima. Larsen, Clark Spencer 2015 Bioarchaeology: Interpreting Behavior from the Human Skeleton. 2nd ed. Cambridge University Press, Cambridge, UK. Las Casas, Bartolome de 1967 [1550] Apologetica Historica Sumaria. Instituto de Investigaciones Historicas, Universidad Nacional Autonoma de Mexico, Mexico City. Latour, Bruno B. 2005 Reassembling the Social: An Introduction to Actor-Network Theory. Oxford University Press, Oxford, UK.

4 1 6   Reference List Lau, George F. 2004 Object of Contention: An Examination of Recuay-Moche Combat Imagery. Cambridge Archaeological Journal 14:163–184. 2011 Andean Expressions: Art and Archaeology of the Recuay Culture. The Iowa Series in Andean Studies. University of Iowa Press, Iowa City. Lewis, Cecil M., Raul Y. Tito, Beatriz Lizarraga, and Anne C. Stone 2004 Land, Language, and Loci: mtDNA in Native Americans and the Genetic History of Peru. American Journal of Physical Anthropology 127:351–360. Lienhardt, R. Godfrey 1961 Divinity and Experience: The Religion of the Dinka. Clarendon Press, Oxford, UK. Longin, R. 1971 New Method of Collagen Extraction for Radiocarbon Dating. Nature 230:241–242. López Austin, Alfredo 1989 Cuerpo Humano e Ideología. Las Concepciones de los Antiguos Nahuas, Universidad Nacional Autónoma de México, Mexico City. López Luján, Leonardo, and Guilhem Oliver (editors) 2010 El Sacrifico Humano en la Tradición Religiosa Mesoamericana. Instituto Nacional de Antropología e Historia/Universidad Nacional Autónoma de México, Mexico City. Losey, Robert J., Sandra Garvie-Lok, Jennifer A. Leonard, M. Anne Katzenberg, Mietje Germonpré, Tatiana Nomokonova, Mikhail V. Sablin, Olga I. Goriunova, Natalia E. Berdnikova, and Nikolai A. Savel’ev 2013 Burying Dogs in Ancient Cis-Baikal, Siberia: Temporal Trends and Relationships with Human Diet and Subsistence Practices. PLoS ONE 8(5):e63740. Lothrop, Samuel K. 1941 Gold Ornaments of Chavín Style from Chongoyape, Peru. American Antiquity 6:250–262. Lovejoy, C. Owen, Richard S. Meindl, Thomas R. Pryzbeck, and Robert P. Mensforth 1985 Chronological Metamorphosis of the Auricular Surface of the Ilium: A New Method for the Determination of Adult Skeletal Age at Death. American Journal of Physical Anthropology 68:14–28. Lovell, Nancy C. 1997 Trauma Analysis in Paleopathology. Yearbook of Physical Anthropology 40:139–170. 2008 Analysis and Interpretation of Skeletal Trauma. In Biological Anthropology of the Human Skeleton, 2nd ed., edited by M. Anne Katzenberg and Shelly Saunders, pp. 341–386. Wiley-Liss, New York. Lozada, Maria Cecilia, and Jane E. Buikstra 2005 Pescadores and Labradores among the Señorio of Chiribaya in Southern Peru. In Us and Them: Archaeology and Ethnicity in the Andes, edited by Richard Martin Reycraft, pp. 206–225. Cotsen Institute of Archaeology Press, University of California, Los Angeles.

Reference List  417 Lucas, Gavin, and Thomas McGovern 2007 Bloody Slaughter: Ritual Decapitation and Display at the Viking Settlement of Hofstaðir, Iceland. European Journal of Archaeology 10:7–30. Luke, James L. 1967 Strangulation as a Method of Homicide. Archives of Pathology 83:64–70. Lutz, Catherine 2007 Militarization. In A Companion to the Anthropology of Politics, edited by David Nugent and Joan Vincent, pp. 318–331. Blackwell, Boston. Lyman R. Lee 1994 Vertebrate Taphonomy. Cambridge University Press, Cambridge, UK. Mackey, Carol J. 2006 Elite Residences at Farfán: A Comparison of the Chimú and Inka Occupations. In Palaces and Power in the Americas: From Peru to the Northwest Coast, edited by Jessica Joyce Christie and Patricia Joan Sarro, pp. 313–352. University of Texas Press, Austin. 2010 The Socioeconomic and Ideological Transformation of Farfán under Inka Rule. In Distant Provinces in the Inka Empire: Towards a Deeper Understanding of Inka Imperialism, edited by Michael Malpass and Sonia Alconini, pp. 221–259. University of Iowa Press, Iowa City. 2011 Chimu Statecraft in the Provinces. In Andean Civilization: A Tribute to Michael Moseley, edited by Joyce Marcus and Patrick Ryan Williams, pp. 325–349. UCLA Cotsen Institute of Archaeology Press, Los Angeles. Mackey, Carol J., and A. M. Ulana Klymyshyn 1990 The Southern Frontier of the Chimu Empire. In The Northern Dynasties: Kingship and Statecraft in Chimor, edited by Michale E. Moseley and Alana Cordy-Collins, pp. 195–226. Dumbarton Oaks, Washington, DC. Mackey, Carol J., and Joanne Pillsbury 2013 Cosmology and Ritual on a Lambayeque Beaker. In Pre-Columbian Art and Archaeology: Essays in Honor of Frederick R. Mayer, edited by Margaret Young-Sánchez, pp. 115–141. Mayer Center for Pre-Columbian and Spanish Colonial Art at the Denver Art Museum, Denver. Marcus, Joyce 2007 Rethinking Ritual. In The Archaeology of Ritual, edited by Evangelos Kyriakidis, pp. 43–76. UCLA Cotsen Institute of Archaeology Press, Los Angeles. Marcus, Joyce, and K. V. Flannery 1994 Ancient Zapotec Ritual and Religion: An Application of the Direct Historical Approach. In The Ancient Mind: Elements of Cognitive Archaeology, edited by Colin Renfrew and Ezra B. W. Zubrow, pp. 55–74. New Directions in Archaeology. Cambridge University Press, Cambridge, UK. Marean, Curtis W., Yoshiko Abe, Peter J. Nilssen, and Elizabeth C. Stone 2001 Estimating the Minimum Number of Skeletal Elements (MNE) in Zooarchaeology: A Review and a New Image-Analysis GIS Approach. American Antiquity 66:333–348. Margerison, Beverley J., and Christopher J. Knüsel 2002 Paleodemographic Comparison of a Catastrophic and an Attrition-

4 1 8   Reference List al Death Assemblage. American Journal of Physical Anthropology 119:134–143. Marshall, L. G. 1989 Bone Modification and the “Laws of Burial.” In Bone Modification, edited by Robsen Bonnichsen and Marcella Sorg, pp. 7–24. Center for the Study of the First Americans, Institute for Quaternary Studies, University of Maine, Orono. Martin, Deborah L., Ryan P. Harrod, and Ventura Pérez (editors) 2012 The Bioarchaeology of Violence. University Press of Florida, Gainesville. Martínez, Juan 1997 Proyecto de Rescate Arqueológico Illimo-Depolti. Technical report on file at the Museo Nacional de Arqueología y Etnografía Hans Henrich Brüning, Lambayeque, Peru. 2011 Proyecto Arqueológico La Joroto-Jayanca-Lambayeque: Informe Temporada 2010. Technical report on file at the Museo Nacional de Arqueología y Etnografía Hans Henrich Brüning, Lambayeque, Perú. Massey, Virginia K., and D. Gentry Steele 1997 A Maya Skull Pit from the Terminal Classic Period, Colha, Belize. In Bones of The Maya: Studies of Ancient Skeletons, edited by Stephen L. Whittington and David M. Reed, pp. 62–77. Smithsonian Institution Press, Washington, DC. Masur, L. J., J.-F. Millaire, and M. Blake Submitted. Peanuts and Prestige in the Andes: Spatial Analysis of Botanical Remains from the Virú Valley. Journal of Ethnobiology. Matos, Eduardo 1986 Muerte al filo de obsidiana. Lecturas Mexicanas, SEP, Mexico City. Matsumoto, Go 2014 El Culto a los Ancestros: Aproximación y Evidencia. In Cultura Sicán: Esplendor Preinca de la Costa Norte, edited by Izumi Shimada, pp. 195–215. Fondo Editorial del Congreso de Perú, Lima. Mauss, Marcel, and Henri Hubert 1899 Essai sur la nature et la fonction du sacrifice. In L’Année Sociologique, Vol. II, pp. 29–38. F. Alcan, Paris. McAnany, Patricia 1995 Living with the Ancestors: Kinship and Kingship in Ancient Maya Society. University of Texas Press, Austin. McAnany, Patricia, Rebecca Storey, and Angela K. Lockard 1999 Mortuary Ritual and Family Politics at Formative and Early Classic K’axob, Belize. Ancient Mesoamerica 10:129–146. McClelland, Donna 2008 Ulluchu—An Elusive Fruit. In The Art and Archaeology of the Moche: An Ancient Andean Society of the Peruvian North Coast, edited by Steve Bourget and Kimberly L. Jones, pp. 43–65. University of Texas Press, Austin. McClymond, Kathryn 2008 Beyond Sacred Violence: A Comparative Study of Sacrifice. Johns Hopkins University Press, Baltimore.

Reference List  4 1 9 McCorkle, Constance M. 1987 Punas, Pastures, and Fields: Grazing Strategies and the Agropastoral Dialectic in an Indigenous Andean Community. In Arid Land Use Strategies and Risk Management in the Andes: A Regional Anthropological Perspective, edited by David L. Browman, pp. 57–80. Westview, Boulder, Colorado. McEwan, Colin, and Martín Van de Guchte 1992 Ancestral Time and Sacred Space in Inca State Ritual. In The Ancient Americas: Art from Sacred Landscapes, edited by Richard F. Townsend, pp. 359–371. Art Institute of Chicago, Chicago. Melbye, Jerry, and Scott I. Fairgrieve 1990 Appendix IV: The Human Remains. In Archaeological Investigations at Saunaktuk, edited by Charles D. Arnold, pp. 85–113. Princes of Wales Northern Heritage Centre, Department of Culture and Communications, Government of the Northwest Territories, Yellowknife, Canada. 1994 A Massacre and Possible Cannibalism in the Canadian Arctic: New Evidence from the Saunaktuk Site (NgTn-1). Arctic Anthropology 31:57–77. Merbs, Charles 2002 Asymmetrical Spondylolysis. American Journal of Physical Anthropology 119:156–174. Meskell, Lynne 1999 Writing the Body in Archaeology. In Reading the Body: Representations and Remains in the Archaeological Record, edited by A. E. Rautman, pp. 13–21. University of Pennsylvania Press, Philadelphia. Meyer-Orlac, Renate 1997 Zur Problematik der “Sonderbestattungen” in der Archaologie. In Sonderbestattungen in der Bronzezeit im Ostlichen Mitteleuropa, edited by Karl-Friedrich Rittershofer, pp. 1–10. Verlag Marie Leidorf GmbH, Espelkamp, Germany. Middleton, John 1987 Lugbara Religion: Ritual and Authority among an East African People. Smithsonian Institution Press, Washington, DC. Millaire, Jean-François 2002 Moche Burial Patterns: An Investigation into Prehispanic Social Structure. British Archaeological Report International Series 1066, Oxford, UK. 2004 The Manipulation of Human Remains in Moche Society: Delayed Burials, Grave Reopening, and Secondary Offerings of Human Bones on the Peruvian North Coast. Latin American Antiquity 15:371–389. 2009a Gallinazo and the Tradición Norcosteña. In Gallinazo: An Early Cultural Tradition on the Peruvian North Coast, edited by Jean-François Millaire and Magali Morlion, pp. 1–16. UCLA Cotsen Institute of Archaeology Press, Los Angeles. 2009b Woven Identities in the Virú Valley. In Gallinazo: An Early Cultural Tradition on the Peruvian North Coast, edited by Jean-François Millaire and Magali Morlion, pp. 149–165. UCLA Cotsen Institute of Archaeology Press, Los Angeles.

4 2 0  Reference List 2010a Moche Political Expansionism as Viewed from Virú: Recent Archaeological Work in the Close Periphery of a Hegemonic City-State System. In New Perspectives on Moche Political Organization, edited by Jeffrey Quilter and Luis Jaime Castillo, pp. 223–251. Dumbarton Oaks, Washington, DC. 2010b Primary State Formation in the Virú Valley, North Coast of Peru. Proceedings of the National Academy of Sciences 107:6186–6191. 2015 The Sacred Character of Ruins on the Peruvian North Coast. In Living with the Dead in the Andes, edited by Izumi Shimada and James L. Fitzsimmons, pp. 50–75. University of Arizona Press, Tucson. Millaire, Jean-François, and Edward Eastaugh 2011 Ancient Urban Morphology in the Virú Valley, Peru: Remote Sensing Work at the Gallinazo Group (100 BC–AD 700). Journal of Field Archaeology 36:289–297. Millaire, Jean-François, and Magali Morlion (editors) 2009 Gallinazo: An Early Cultural Tradition on the Peruvian North Coast. UCLA Cotsen Institute of Archaeology Press, Los Angeles. Millaire, Jean-François, and Flannery Surette 2011 Un Fardo Funerario de ca. AD 1150 en la Huaca Santa Clara, Valle de Virú. Bulletin del’Institut Français d’Études Andines 40:289–305. Miller, George R., and Richard L. Burger 1995 Our Father the Cayman, Our Dinner the Llama: Animal Utilization at Chavin de Huantar, Peru. American Antiquity 60(3):421–458. Milner, George, and Virginia Smith 1989 Carnivore Alteration of Human Bone from a Late Prehistoric Site in Illinois. American Journal of Physical Anthropology 79:43–49. Mock, Shirley Boteler 1998 The Sowing and the Dawning: Termination, Dedication, and Transformation in the Archaeological and Ethnographic Record of Mesoamerica. University of New Mexico Press, Albuquerque. Moens, Luc, Alex Von Bohlen, and Peter Vandenabeele 2000 X-Ray Fluorescence. In Modern Analytical Methods in Art and Archaeology, edited by Enrico Ciliberto and Giuseppe Spoto, pp. 55–79. John Wiley and Sons, New York. Mogrovejo, Juan 1995 La Evidencia Funeraria Mochica de Huaca de la Cruz, Valle de Virú. Unpublished Licenciatura thesis, Departamento de Letras y Ciencias Humanas, Pontificia Universidad Católica del Perú, Lima. Molina, Cristóbal de 1943 [1575] Fábulas y Ritos de los Incas. In Las Crónicas de los Molinas. Los Pequeños Grandes Libros de Historia Americana, Lima. Montenegro, Jorge 1993 Estillo Cajamarca Costeño: Una Aproximación. Actas de IX Congreso Peruano del Hombre y la Cultura Andina 9:137–149. Montoya Vera, Maria 1996 Implicaciones del Estudio de Semillas Rituales en la Epoca Prehispánica. Revista del Museo de Arqueología Trujillo 6:203–219.

Reference List  4 2 1 2004 Complejo De Ofrendas Rituales y Su Asociación a Sacrificios Humanos De Niños En La Época Chimú En El Valle De Moche. In Desarrollo Arqueológico de la Costa Norte del Perú, edited by Luis Valle Alvarez, pp. 27–48. SIAN, Trujillo. Moore, Jerry D., and Carol J. Mackey 2008 The Chimú Empire. In Handbook of South American Archaeology, edited by Helaine Silverman and William H. Isbell, pp. 783–807. Springer, New York. Moorrees, C. F. A., E. A. Fanning, and E. E. Hunt 1963 Age Formation by Stages for Ten Permanent Teeth. Journal of Dental Anthropology 42:1490–1502. Morlan, Richard E. 1984 Toward the Definition of Criteria for the Recognition of Artificial Bone Alterations. Quaternary Research 22:160–171. Morlion, Magali 1999 La Stratification du Rituel Religieux à la Période Préclassique en Mésoamérique. Unpublished MSc thesis, Université de Montréal, Montreal, Canada. Morris, Craig, and Adriana von Hagen 2011 The Incas: Lords of the Four Quarters. Thames and Hudson, New York. Morris, Ellen F. 2007 Sacrifice for the State: First Dynasty Royal Funerals and the Rites at Macramallah’s Rectangle. In Performing Death: Social Analyses of Funerary Traditions in the Ancient Near East and Mediterranean, edited by Nicola Laneri, pp. 15–38. Oriental Institute of the University of Chicago, Chicago. Moseley, Michael E. 1992 The Incas and their Ancestors. Thames and Hudson, London. 2001 The Incas and their Ancestors. 2nd ed. Thames and Hudson, London. Moseley, Michael E., and Alana Cordy-Collins (editors) 1990 The Northern Dynasties: Kingship and Statecraft in Chimor. Dumbarton Oaks, Washington DC. Moseley, Michael E., and Kent Day (editors) 1982 Chan Chan: Andean Desert City. University of New Mexico Press, Albuquerque. Moseley, Michael E., Richard A. Feldman, Charles R. Ortloff, and Alfredo Narváez 1983 Principles of Agrarian Collapse in the Cordillera Negra, Peru. Annals of Carnegie Museum 52:299–327. Moser, Christopher L. 1973 Human Decapitation in Ancient Mesoamerica. Dumbarton Oaks, Washington, DC. 1974 Ritual Decapitation in Moche Art. Archaeology 27:30–37. Mujica, Elias (editor) 2007 El Brujo: Huaca Cao, Centro Ceremonial Moche en el Valle de Chicama/ Huaca Cao, a Moche Ceremonial Center in the Chicama Valley. Fundación Wiese, Lima. Mundorff, Amy, Robert Shaler, Erik Bieschke, and Elaine Mar-Cash 2008 Marrying Anthropology and DNA: Essential for Solving Complex

4 2 2   Reference List Commingling Problems in Cases of Extreme Fragmentation. In Recovery, Analysis, and Identification of Commingled Human Remains, edited by Bradley Adams and John Byrd, pp. 285–300. Humana Press, New York. Muno, Sarah K. 2014 La Sociedad Sicán: Una Visión Desde los Esqueletos. In Cultura Sicán: Esplendor Preinca de la Costa Norte, edited by Izumi Shimada. Fondo Editorial del Congreso de Perú, Lima. Murad, Turhon, and Margie Boddy 1987 A Case with Bear Facts. Journal of Forensic Sciences 32:1819–1826. Murphy, Brett P., and David M. J. S. Bowman 2006 Kangaroo Metabolism does not Cause the Relationship between Bone Collagen δ15N and Water Availability. Functional Ecology 20:1062–1069. Murphy, Eileen (editor) 2008 Deviant Burial in the Archaeological Record. Oxbow, Oxford, UK. Murphy, Melissa S. 2004 From Bare Bones to Mummified: Understanding Health and Disease in an Inca Community. Unpublished PhD dissertation, Department of Anthropology, University of Pennsylvania, Philadelphia. Murphy, Melissa Scott, Catherine M. Gaither, Elena Goycochea, John W. Verano, and Guillermo A. Cock 2010 Violence and Weapon-Related Trauma at Puruchuco-Huaquerones, Peru. American Journal of Physical Anthropology 142:636–649. Murra, John V. 1962 Cloth and Its Functions in the Inca State. American Anthropologist 64:710–728. 1965 Herds and Herders in the Inca State. In Man, Culture, and Animals: The Role of Animals in Human Ecological Adjustments, edited by A. Leeds and A. P. Vayda, pp. 85–215. American Association for the Advancement of Science, Washington, DC. 1980 The Economic Organization of the Inka State. JAI Press, Greenwich, Connecticut. Nadasdy, Paul 2007 The Gift in the Animal: The Ontology of Hunting and Human–Animal Sociality. American Ethnologist 34:25–43. Nájera, Martha Ilia 1987 El Don de la Sangre en el Equilibrio Cósmico. Centro de Estudios Mayas, Universidad Nacional Autónoma de México, Mexico City. Narváez, Alfredo 1995a Death in Ancient Túcume. The Southern Cemetery and Huaca Facho. In Pyramids of Túcume: The Quest for Peru’s Forgotten City, by Thor Heyerdahl, Daniel H. Sandweiss, and Alfredo Narváez, pp. 169–178. Thames and Hudson, London. 1995b The Pyramids of Túcume: The Monumental Sector. In Pyramids of Túcume: The Quest for Peru’s Forgotten City, by Thor Heyerdahl, Daniel H. Sandweiss, and Alfredo Narváez, pp. 79–130. Thames and Hudson, London.

Reference List  4 23 Narváez Vargas, Alfredo, and Bernarda Delgado Elias (editors) 2011 Huaca Las Balsas de Túcume: Arte Mural Lambayeque. Editoral Súper Gráfica, Lima. In press Huaca I De Túcume. Museo de Sitio de Túcume, Túcume. Nawrocki, Stephen P., John E. Pless, Dean A. Hawley, and Scott A. Wagner 1997 Fluvial Transport of Human Crania. In Forensic Taphonomy: The Postmortem Fate of Human Remains, edited by William D. Haglund and Marcella H. Sorg, pp. 529–552. CRC Press, Boca Raton, Florida. Needham, R. 1967 Percussion and Transition. Man 2:606–614. Nelson, Andrew J. 1998 Wandering Bones: Archaeology, Forensic Science, and Moche Burial Practices. International Journal of Osteoarchaeology 8:192–212. Nelson, Andrew J., and Luis Jaime Castillo 1997 Huesos a la Deriva: Tafonomía y Tratamiento Funerario en Entierros Mochica Tardío de San José de Moro. Boletín de Arqueologia PUCP 1: 137– 163. Pontificia Universidad Católica del Perú, Lima. Nelson, Andrew J., and Carol J. Mackey 2011 A Reanalysis of the Avispas Burial Mound at Chan Chan. Paper presented at the 76th Annual Meeting of the Society for American Archaeology, Sacramento, California. Nelson, Andrew J., Christine S. Nelson, Luis Jaime Castillo, and Carol J. Mackey 2000 La Osteobiografia De Una Hilandera Precolumbina: La Mujer Detras De La Mascara. Iconos 4(2):30–43. Nguyen, Vinh-Kim, and Karine Peschard 2003 Anthropology, Inequality, and Disease: A Review. Annual Review of Anthropology 32:447–474. Nielsen, Axel M. 2001 Ethnoarchaeological Perspectives on Caravan Trade in the South-Central Andes. In Ethnoarchaeology of Andean South America: Contributions to Archaeological Method and Theory, edited by Lawrence A. Kuznar, pp. 163–201. International Monographs in Prehistory. Ethnoarchaeological Series 4, Ann Arbor, Michigan. Nordstrom, Carolyn 1998 Deadly Myths of Aggression. Aggressive Behavior 24:147–159. Nystrom, Kenneth C. 2006 Late Chachapoya Population Structure Prior to Inka Conquest. American Journal of Physical Anthropology 131:334–342. Ochatoma, José A., and Martha R. Cabrera 2002 Religious Ideology and Military Organization in the Iconography of a D-shaped Ceremonial Precinct at Conchopata. In Andean Archaeology II: Art, Landscape, and Society, edited by Helaine Silverman and William H. Isbell, pp. 225–247. Kluwer, New York. O’Connell, Timothy C., and Richard E. M. Hedges 1999 Investigations into the Effect of Diet on Modern Human Hair Isotopic Values. American Journal of Physical Anthropology 108:409–425.

4 2 4   Reference List Oestigaard, Terje 2000 Sacrifices of Raw, Cooked, and Burnt Humans. Norwegian Archaeological Review 33:41–58. Ogburn, Dennis E. 2007 Human Trophies in the Late Pre-Hispanic Andes. In The Taking and Displaying of Human Body Parts as Trophies by Amerindians, edited by Richard J. Chacon and David H. Dye, pp. 505–522. Springer, New York. O’Leary, Marion H. 1981 Carbon Isotope Fractionation in Plants. Phytochemistry 20:553–567. Olsen, Sandra L., and Pat Shipman 1988 Surface Modification on Bone: Trampling versus Butchery. Journal of Archaeological Science 15:535–553. 1994 Cutmarks and Perimortem Treatment of Skeletal Remains on the Northern Plains. In Skeletal Biology in the Great Plains: A Multidisciplinary View, edited by Douglas Owsley and Richard Jantz, pp. 377–387. Smithsonian Institution Press, Washington, DC. Onuki, Yoshio 1993 Las Actividades Cermoniales Tempranas en la Cuenca del Alto Huallaga y Algunos Problemas Generales. In El Mundo Ceremonial Andino, edited by Luis Millones and Yoshio Onuki, pp. 69–96. Senri Ethnological Studies No. 37. National Ethnographic Museum, Osaka. Onuki, Yoshio, and Kinya Inokuchi 2011 Gemelos Prístinos: El Tesoro del Templo Kuntur Wasi. Fondo Editorial del Congreso del Perú and Minera Yanacocha, Lima. Orchard, T. J. 2005 The Use of Statistical Size Estimations in Minimum Number Calculations. International Journal of Osteoarchaeology 15:351–359. Orlove, Benjamin S. 1977 Alpacas, Sheep, and Men: The Wool Export Economy and Regional Society of Southern Peru. Academic Press, New York. Ortner, Donald J. 2003 Identification of Pathological Conditions in Human Skeletal Remains. 2nd ed. Academic Press, Boston. Ortner, Donald J., David W. Von Endt, and Mary S. Robinson 1972 The Effect of Temperature on Protein Decay in Bone: Its Significance in Nitrogen Dating of Archaeological Specimens. American Antiquity 37:514–520. Osborne, Robin 2004 Hoards, Votives, Offerings: The Archaeology of the Dedicated Object. World Archaeology 36:1–10. Otterbein, Keith F. 2000 Killing of Captured Enemies: A Cross-cultural Study. Current Anthropology 41:439–443. Otto, Ton, Henrik Thrane, and Helle Vandkilde (editors) 2006 Warfare and Society: Archaeological and Social Anthropological Perspectives. Aarhus University Press, Aarhus.

Reference List  4 2 5 Parker Pearson, Mike 2000 The Archaeology of Death and Burial. Texas A&M Press, College Station. Pease, Franklin G. Y. 1982 The Formation of Tawantinsuyu: Mechanisms of Colonization and Relationship with Ethnic Groups. In The Inca and Aztec States, 1400–1800: Anthropology and History, edited by George A. Collier, Ronato I. Rosaldo, and John D. Wirth, pp. 173–198. Academic Press, New York. Pechenkina, Ekaterina A., and Mercedes Delgado 2006 Dimensions of Health and Social Structure in the Early Intermediate Period Cemetery at Villa El Salvador, Peru. American Journal of Physical Anthropology 131:218–235. Pereira, Grégory 1999 Potrero de Guadalupe: anthropologie funéraire d’une communauté pré-tarasque du nord du Michoacán, Mexique. B.A.R. International Series 816, Oxford, UK. 2005 The Utilization of Grooved Human Bones: A Reanalysis of Artificially Modified Human Bones Excavated by Carl Lumholtz at Zacapu, Michoacán, Mexico. Latin American Antiquity 16:295–312. Pereira, Grégory, Saburo Sugiyama, and Michael Spence 2006 Regroupement des morts, monumentalité et sacrifice: l’ exemple de Teotihuacan (Mexique). In Aux origines du regroupement des morts: du fait singulier à la coutume, edited by D. Castex, P. Coutaud, H. Duday, F. Le Mort, and A.-M. Tillier, pp. 249–280. Maison des Sciences de l’Homme d’Aquitaine, Bordeaux. Phenice, T. W. 1969 A Newly Developed Visual Method of Sexing in the Os Pubis. American Journal of Physical Anthropology 30:297–301. Phillips, Sara S. 2009 Warriors, Victims, and the Merely Accident Prone: Fracture Patterns in Moche Skeletal Remains from Northern Coastal Peru. Unpublished PhD dissertation, Department of Anthropology, Tulane University, New Orleans. Phillips, Sara S., and John W. Verano 2005 Trauma Patterns in the Massacre Victims from Punta Lobos, Northern Coastal Peru. Paper presented at the 74th Annual Meeting of the American Association of Physical Anthropologists, Milwaukee, Wisconsin. Pierce, Mary Clyde, Gina E. Bertocci, Eva Vogeley, and Morey S. Moreland 2004 Evaluating Long Bone Fractures in Children: A Biomechanical Approach with Illustrative Cases. Child Abuse & Neglect 28:505–524 Pijoan, Carmen, and Xabier Lizarraga (editors) 2004 Perspectiva Tafonómica: Evidencias de Alteraciones en Restos Oseos del México Prehispánico. Instituto Nacional de Antropología e Historia, Mexico City. Pijoan, Carmen, and Josefina Mansilla 1997 Evidence for Human Sacrifice, Bone Modification and Cannibalism in Ancient Mexico. In Troubled Times: Violence and Warfare in the Past,

4 2 6   Reference List edited by Debra L. Martin and David W. Frayer, pp. 217–239. Gordon and Breach, Amsterdam. Pijoan, Carmen María, and Alejandro Pastrana 1987 Método Para Registrado de Marcas de Corte en Huesos Humanos: El Caso de Tlatelcomila, Tetelpan, D.F. Estudios de Antropología Biológica 3:561–583. Pijoan, Carmen María, Alejandro Pastrana, and Consuelo Maquivar 1989 El Tzompantli de Tlatelolco: Una Evidencia de Sacrificio Humano. Estudios de Antropología Biológica 4:561–583. Pillsbury, Joanne 1996 The Thorny Oyster and the Origins of Empire: Implications of Recently Uncovered Spondylus Imagery from Chan Chan, Peru. Latin American Antiquity 7:313–340. Pillsbury, Joanne (editor) 2001 Moche Art and Archaeology in Ancient Peru. National Gallery of Art, Washington, DC. Pizzaro, Pedro 1921 [1571] Relation of the Discovery and Conquest of the Kingdoms of Peru. Vol. 1. Translation by Phillip Means. New York: Cortes Society. Pluskowski, Aleksander (editor) 2012 The Ritual Killing and Burial of Animals: European Perspectives. Oxbow, Oxford, UK. Pokines, James, and Steven Symes 2012 Manual of Forensic Taphonomy. CRC Press, Boca Raton, Florida. Polo de Ondegardo, Juan 1571 Relación de los Fundamentos acerca del Notable Daño que Resulta de no Guarar a los Indios sus Fueros. Colección de Libros y Documentos referentes a la Historia del Perú 1:45–188. Polson, Cyril J., and David J. Gee 1973 The Essentials of Forensic Medicine. 3rd ed. Pergamon Press, Oxford, UK. Porter, Anne, and Glenn M. Schwartz (editors) 2012 Sacred Killing: The Archaeology of Sacrifice in the Ancient Near East. Eisenbrauns, Winona Lake, Indiana. Porter, R. E. 1989 Normal Development of Movement and Function: Child and Adolescent. In Physical Therapy, edited by Rosemary M. Scully, and Marylou R. Barnes, pp. 83–98. JB Lippincott, Philadelphia. Pozorski, Shelia G. 1976 Prehistoric Subsistence Patterns and Site Economics in the Moche Valley, Peru. Unpublished PhD dissertation, Department of Anthropology, University of Texas, Austin. 1979 Prehistoric Diet and Subsistence of the Moche Valley, Peru. World Archaeology 11:163–184. Pozorski, Shelia G., and Thomas Pozorski 1992 Early Civilization in the Casma Valley, Peru. Antiquity 66:845–870. 2002 The Sechín Alto Complex and Its Place within Casma Calley Initial Period Development. Andean Archaeology I: Variations in Sociopolitical

Reference List  4 27 Organization, edited by William H. Isbell and Helaine Silverman, pp. 21–51. Kluwer, New York. 2008 Early Cultural Complexity on the Coast of Peru. In Handbook of South American Archaeology, edited by Helaine Silverman and William H. Isbell, pp. 607–631. Springer, New York. Pozorski, Thomas 1971 Survey and Excavations of Burial Platforms at Chan Chan, Peru. Unpublished Senior Honors Thesis, Department of Anthropology, Harvard University, Cambridge, Massachusetts. Prieto, O. Gabriel 2010 Aproximaciones a la Configuración Política Lambayeque: Una Perspectiva desde el Sitio de San Jose de Moro, Valle de Jequetepeque. In Perspectivas Comparativas sobre la Arqueología de la Costa Sudamericana/ Comparative Perspectives on the Archaeology of Coastal South America, edited by Robyn E. Cutright, Enrique López-Hurtado, and Alexander J. Martin, pp. 232–246. Fondo Editorial Pontificia Universidad Católica del Peru, Lima. Prieto O. Gabriel, Nicolás Goepfert, and Katya Valladares 2015 Sacrificios de Niños, Adolescentes, y Camélidos Jóvenes durante el Intermedio Tardío en la Periferia de Chan Chan, Valle de Moche, Costa Norte del Perú. Arqueología y Sociedad, Museo de Arqueología y Antropología de San Marcos. 27:255–296. Proulx, Donald A. 1982 Territoriality in the Early Intermediate Period: The Case of Moche and Recuay. Ñawpa Pacha 20:83–96. 1999 Nasca Headhunting and the Ritual Use of Trophy Heads. In Nasca: Geheimnisvolle Zeichen im Alten Peru, edited by Judith Rickenbach, pp. 79–87. Museum Rietberg Zürich, Zürich. 2001 Ritual Uses of Trophy Heads in Ancient Nasca Pottery. In Ritual Sacrifice in Ancient Peru, edited by Elizabeth P. Benson and Anita G. Cook, pp. 119–136. University of Texas Press, Austin. 2004 Pañamarca and the Moche Presence in the Nepeña Valley Revisited. Paper presented at the 64th Annual Meeting of the Society for American Archaeology, Montréal, Québec, Canada. Purdue, Basil N. 2000 Asphyxial and Related Deaths. In The Pathology of Trauma, edited by J. K. Mason and B. N. Purdue, pp. 230–252. Arnold Publishing, London. Purin, Sergio 1990 Inca—Perú: 3000 Ans d’Histoire. Musées Royaux d’Art et d’Histoire, Brussels. Püschel, Klaus, Elisabeth Turk, and Holger Lach 2004 Asphyxia-Related Deaths. Forensic Science International 144:211–214. Quilter, Jeffrey 1985 Architecture and Chronology at El Paraíso, Peru. Journal of Field Archaeology 12:279–297. 1989 Life and Death at Paloma: Society and Mortuary Practices in a Preceramic Peruvian Village. University of Iowa Press, Iowa City.

4 2 8   Reference List 1997

The Narrative Approach to Moche Iconography. Latin American Antiquity 8:113–133. 2001 Moche Mimesis: Continuity and Change in Public Art in Early Peru. In Moche Art and Archaeology in Ancient Peru, edited by Joanne Pillsbury, pp. 21–45. National Gallery of Art, Washington, DC. 2002 Moche Politics, Religion, and Warfare. Journal of World Prehistory 16:145–195. 2008 Art and Moche Martial Arts. In The Art and Archaeology of the Moche: An Ancient Andean Society of the Peruvian North Coast, edited by Steve Bourget and Kimberly L. Jones, pp. 215–228. University of Texas Press, Austin. Quilter, Jeffery, and Luis Jaime Castillo B. (editors) 2010 New Perspectives on Moche Political Organization. Dumbarton Oaks: Washington, DC. Quilter, Jeffrey, and Michele L. Koons 2012 The Fall of the Moche: A Critique of Claims for South America’s First State. Latin American Antiquity 23:127–143. Raemsch, Carol A. 1993 Mechanical Procedures Involved in Bone Dismemberment and Defleshing in Prehistoric Michigan. Midcontinental Journal of Archaeology 18:217–244. Rahimi, Babak 2004 The Rebound Theater State: The Politics of the Safavid Camel Sacrifice Rituals, 1598–1695 C.E. Iranian Studies 37:451–478. Rakita, Gordon F. M., Jane E. Buikstra, Lane Anderson Beck, and Sloan R. Williams (editors) 2005 Interacting with the Dead: Perspectives on Mortuary Archaeology for the New Millennium. University Press of Florida, Gainesville. Ramírez, Susan E. 1990 The Inca Conquest of the North Coast: A Historian’s View. In The Northern Dynasties: Kingship and Statecraft in Chimor, edited by Michael Moseley and Alana Cordy-Collins, pp. 507–537. Dumbarton Oaks, Washington DC. 1996 The World Upside Down: Cross-Cultural Contact and Conflict in SixteenthCentury Peru. Stanford University Press, Stanford. 1998 Rich Man, Poor Man, Beggar Man, or Chief: Two Views of the Concept of Wealth in 16th Century Peru. In Dead Giveaways: Indigenous Testaments of Colonial Spanish America, edited by Susan Kellogg and Matthew Restall, pp. 215–248. University of Utah Press, Salt Lake City. 2004 Comments on “An Integrated Analysis of Pre-Hispanic Mortuary Practices.” Current Anthropology 45:394–395. 2005 To Feed and Be Fed: The Cosmological Bases of Authority and Identity in the Andes. Stanford University Press, Stanford. Rang, Mercer, and Dennis Ray Wenger 2006 Children Are Not Just Small Adults. In Rang’s Children’s Fractures. 3rd ed., edited by Mercer Rang, Maya E. Pring, and Dennis Ray Wenger, pp. 1–25. Lippincott Williams and Wilkins, Philadelphia.

Reference List  4 2 9 Rao, Dinesh 2010 Ligature Strangulation. In Forensic Pathology online. Available at: www .forensicpathologyonline.com. Accessed January 29, 2012. Rappaport, Roy A. 2004 Ritual and Religion in the Making of Humanity. Cambridge University Press, Cambridge, UK. Rasmussen, Susan J. 2002 Animal Sacrifice and the Problem of Translation: The Construction of Meaning in Tuareg Sacrifice. Journal of Ritual Studies 16:141–164. Ravines, Rogger, César Alva Velásquez, Proyecto de Rescate Arqueológico Jequetepeque, Instituto Nacional de Cultura (Peru), and Proyecto de Irrigación Jequetepeque-Zaña: Dirección Ejecutiva 1982 Arqueología del Valle Medio del Jequetepeque. Proyecto de Rescate Arqueológico Jequetepeque, Lima. Rea, Amadeo M. 1986 Black Vultures and Human Victims: Archaeological Evidence from Pacatnamu. In The Pacatnamu Papers: Volume 1, edited by Chrisopher B. Donnan and Guillermo A. Cock, pp. 139–144. Museum of Cultural History, Los Angeles. Reimer, P. J., M. G. L. Baillie, E. Bard, E. Bayliss, J. W. Beck, P. G. Blackwell, C. Bronk Ramsey, C. E. Buck, G. S. Burr, R. L. Edwards, M. Friedrich, P. M. Grootes, T. P. Guilderson, Irka Hajdas, Timothy J. Heaton, A. G. Hogg, K. A. Hughen, K. F. Kaiser, B. Kromer, F. G. McCormac, S. W. Manning, R. W. Reimer, D. A. Richards, John R. Southon, S. Talamo, C. S. M. Turney, J. van der Plicht, and C. E. Weyhenmeyer 2009 IntCal09 and Marine09 Radiocarbon Age Calibration Curves, 0–50,000 Years cal BP. Radiocarbon 51:1111–1150. Reinhard, Johan 1996 Peru’s Ice Maidens. National Geographic 189:62–81. 1997 Mummies of Peru. National Geographic 191:36–43. 1998 Discovering the Inca Ice Maiden: My Adventures on Ampato. National Geographic Society, Washington, DC. 2005 The Ice Maiden: Inca Mummies, Mountain Gods, and Sacred Sites in the Andes. National Geographic Society, Washington, DC. Reinhard, Johan, and María Constanza Ceruti 2000 Investigaciones Arqueológicas en el Volcán Llullaillaco: Complejo Ceremonial Incaico de Alta Montaña. EUCASA-Ediciones Universidad Católica de Salta, Salta, Argentina. 2010 Inca Rituals and Sacred Mountains: A Study of the World’s Highest Archaeological Sites. UCLA Cotsen Institute of Archaeology Press, Los Angeles. Rick, John 2013 Canals, Sacrifice, and Water Ritualism at Chavín de Huantar, Peru. Paper presented at the 78th Annual Meeting of the Society for American Archaeology, Honolulu, Hawaii. Robb, John 1997 Violence and Gender in Early Italy. In Troubled Times: Violence and

4 3 0  Reference List Warfare in the Past, edited by Deborah L. Martin and David W. Frayer, pp. 111–144. Gordon and Breach, Amsterdam. Robb, John, Renzo Bigazzi, Luca Lazzarini, Caterina Scarsini, and Fiorenza Sonego 2001 Social “Status” and Biological “Status”: A Comparison of Grave Goods and Skeletal Indicators from Pontecagnano. American Journal of Physical Anthropology 115:213–222. Rockwood, Charles A., and David P. Green 1984 Fractures in Adults. J. B. Lippincott, Philadelphia. Rodríguez, Julio 1995 Informe Técnico de las Excavaciones de Rescate Arqueológico en el Monticulo I de la Caleta de San José (Temporada 6 Marzo al 6 de Mayo 1995. Technical report on file at the Museo Nacional de Arqueología y Etnografía Hans Henrich Brüning, Lambayeque, Peru. Rodriguez, William C., and William M. Bass 1983 Insect Activity and Its Relationship to Decay Rates of Human Cadavers in East Tennessee. Journal of Forensic Sciences 28:423–432. Roe, Peter G. 2008 How to Build a Raptor: Why the Dumbarton Oaks “Scaled Cayman” Callango Textile Is Really a Chavín Jaguaroid Harpy Eagle. In Chavín: Art, Architecture, and Culture, edited by William J. Conklin and Jeffrey Quilter, pp. 181–216. UCLA Cotsen Institute of Archaeology Press, Los Angeles. Rosales Tham, Teresa E. 1999 Informe Final: Proyecto Manejo Ecosustentable y Desarrollo Cultural del Complejo Arqueológico Santa Rita “B.” Temporada 1998. Report on file with the Instituto Nacional de Cultura, Lima. Rossen, Jack 2011 Las Pircas Phase (9800–7800 BP). In From Foraging to Farming in the Andes: New Perspectives on Food Production and Social Organization, edited by Tom D. Dillehay, pp. 95–115. Cambridge University Press, Cambridge, UK. Rossen, Robert, Herman Kabat, and John. P. Anderson 1943 Acute Arrest of Cerebral Circulation in Man. Archives of Neurology and Psychiatry 50:510–528. Rostworowski de Diez Canseco, Maria 1999 History of the Inca Realm. Cambridge University Press, Cambridge, UK. Rousseeuw, Peter J. 1987 Silhouettes: A Graphical Aid in the Interpretation and Validation of Cluster Analysis. Journal of Computational and Applied Mathematics 20:53–65. Rousseeuw, Peter J., A. Struyf, M. Hubert, and M. Maechler 2006 Cluster: Cluster Analysis Extended. R package version 1.11.4. Available at: http://www.R-project.org. Rowe, Anne P. 1996 Chimú Textiles. In Andean Art at Dumbarton Oaks, Volume 2, edited by Elizabeth Hill Boone, pp. 425–436. Dumbarton Oaks, Washington, DC.

Reference List  4 31 Rowe, John H. 1946 Inca Culture at the Time of the Spanish Conquest. In Handbook of South American Indians, Volume 2: The Andean Civilizations, edited by Julian H. Steward, pp. 183–330. Smithsonian Institution, Bureau of American Ethnology, Washington, DC. 1948 The Kingdom of Chimor. Acta Americana 6:26–59. 1982 Inca Policies and Institutions Relating to the Cultural Unification of the Empire. In The Inca and Aztec States, 1400–1800: Anthropology and History, edited by George A. Collier, Ronato I. Rosaldo, and John D. Wirth, pp. 93–118. Academic Press, New York. Rubiños y Andrade, Don Justo Modesto 1936 [1782] Noticia Previa por el Liz. D. Justo Modesto Rubiños, y Andrade, Cura de Mórrope Año de 1782. Revista Historica 10:291–363. Rucabado, Julio 2008 Prácticas Funerarias de Elite en San José de Moro durante la Fase Transicional Temprana: El Caso de la Tumba Colectiva M-U615. In Arqueologia Mochica: Nuevos Enfoques, edited by Luis Jaime Castillo, Helaine Bernier, Julio Rucabado, and Gregory Lockard, pp. 359–380. Fondo Editorial de la Pontificia Universidad Católica del Perú, Lima. Rucabado, Julio, and Luis Jaime Castillo 2003 El Period Transicional en San José de Moro. In Moche Hacia el Final del Milenio, edited by Santiago Uceda and Elias Mujica, pp. 15–42. Universidad Nacional de Trujillo y Pontificia Universidad Católica del Perú, Lima. Ruiz Rosell, Karim 2008 La Tumba M-U1411: Un Entierro Mochica Medio de Elite en el Cementerio de San José de Moro. In Arqueologia Mochica: Nuevos Enfoques, edited by Luis Jaime Castillo, Helaine Bernier, Julio Rucabado, and Gregory Lockard, pp. 381–396. Fondo Editorial de la Pontificia Universidad Católica del Perú, Lima. Russell, Nerissa 2012 Social Zooarchaeology: Humans and Animals in Prehistory. Cambridge University Press, Cambridge, UK. Russell, Nerissa, and Bleda S. During 2006 Worthy Is the Lamb: A Double Burial at Neolithic Çatalhöyük (Turkey). Paléorient 32(1):73–84. Ryan, Kathleen, Karega Munene, Samuel M. Kahinju, and Paul N. Kunoni 2000 Ethnographic Perspectives on Cattle Management in Semi-Arid Environments: A Case Study from Maasailand. In The Origins and Development of African Livestock: Archaeology, Genetics, Linguistics, and Ethnography, edited by Rodger M. Blench and Kevin C. MacDonald, pp. 462–477. UCL Press, London. Sage, Rowan, and Robert Pearcy 2004 The Physiological Ecology of C4 Photosynthesis. In Photosynthesis: Physiology and Metabolism, edited by Richard C. Leegood, Thomas D. Sharkey, and Susanne von Caemmerer, pp. 497–532. Kluwer, New York.

4 3 2   Reference List Salomon, Frank 1995 “The Beautiful Grandparents”: Andean Ancestor Shrines and Mortuary Ritual as Seen through Colonial Records.” In Tombs for the Living: Andean Mortuary Practices, edited by Tom D. Dillehay, pp. 315–353. Dumbarton Oaks, Washington, DC. Samaniego, Lorenzo, Mercedes Cárdenas, Henning Bischoff, Peter Kaulicke, Ermán Guzmán, and Wilder León 1995 Arqueología de Cerro Sechín. Tomo II. Escultura. Pontificia Universidad Católica del Perú, Lima. Sandweiss, Daniel H., and Elizabeth S. Wing 1997 Ritual Rodents: The Guinea Pigs of Chincha, Peru. Journal of Field Archaeology 24:47–58. Sapolsky, Robert M. 2004 Social Status and Health in Humans and Other Animals. Annual Review of Anthropology 33:393–418. Sarmiento de Gamboa, Pedro 1942 Historia de los Incas. Emecé, Buenos Aires, Argentina. Sauer, Norman J. 1984 Manner of Death: Skeletal Evidence of Blunt and Sharp Instrument Wounds. In Human Identification. Case Studies in Forensic Anthropology, edited by Ted A. Rathbun and Jane E. Buikstra, pp. 176–184. Charles C. Thomas, Springfield, Illinois. 1998 The Timing of Injuries and Manner of Death: Distinguishing among Antemortem, Perimortem, and Postmortem Trauma. In Forensic Osteology, edited by Kathleen J. Reichs, pp. 321–332. Charles C. Thomas, Springfield, Illinois. Saul, Frank P., and Julie Mather Saul 1989 Osteobiography: A Maya Example. In Reconstruction of Life from the Skeleton, edited by Mehmet Yasar Iscan and Kenneth A. R. Kennedy, pp. 287– 302. Wiley-Liss, New York. Saunders, Nicholas J. (editor) 1998 Icons of Power: Feline Symbolism in the Americas. Routledge, London. Saxe, Arthur A. 1970 Social Dimensions of Mortuary Practices. Unpublished PhD dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Scarduelli, Pietro 1990 Accumulation of Heads, Distribution of Food: The Image of Power in Nias. Bijdragen tot de Taal-, Land-, en Volkenkunde 146:448–462. Schaedel, Richard P. 1951 Moche Murals at Pañamarca. Archaeology 4:145–154. Schaefer, Maureen C., and Sue M. Black 2007 Epiphyseal Union Sequencing: Aiding in the Recognition and Sorting of Commingled Remains. Journal of Forensic Science 2007:277–285. Scheper-Hughes, Nancy, and Philippe Bourgois (editors) 2004 Violence in War and Peace. Blackwell, Malden, Massachusetts.

Reference List  4 3 3 Scheuer, Louise, and Sue Black 2000 Developmental Juvenile Osteology. Elsevier/Academic Press, Amsterdam. Schindler, Helmut 2000 The Norbert Mayrock Art Collection from Ancient Peru. Staatliches Museum für Völkerkunde München, Munich. Schoeninger, Margaret J., and Michael J. DeNiro 1984 Nitrogen and Carbon Isotopic Composition of Bone Collagen from Marine and Terrestrial Animals. Geochimica et Cosmochimica Acta 48:625–639. Schumacher, A. 1982 On the Significance of Stature in Human Society. Journal of Human Evolution 11:697–701. Schurr, Mark R. 1997 Stable Nitrogen Isotopes as Evidence for the Age of Weaning at the Angel Site: A Comparison of Isotopic and Demographic Measures of Weaning Age. Journal of Archaeological Science 24:919–927. Schwartz, Glenn M. 2012 Archaeology and Sacrifice. In Sacred Killing: The Archaeology of Sacrifice in the Ancient Near East, edited by Anne Porter and Glenn M. Schwartz, pp. 1–32. Eisenbrauns, Winona Lake, Indiana. Scola, Adriane Michelle 2004 Molecular Anthropology and the Punta Lobos Assemblage: DNA-Based Sex Identification of Juveniles from Ancient Hair Samples. Unpublished MA thesis, Department of Anthropology, University of Tennessee Knoxville, Knoxville. Scott, G. Richard, and Christy G. Turner II 1997 The Anthropology of Modern Human Teeth. Cambridge University Press, Cambridge, UK. Segura, Rafael A., and Izumi Shimada 2014 Examinando la Interacción Cultural entre Sicán Medio y la Costa Central, Hacia 1000 d.C. In Cultura Sicán: Esplendor Preinca de la Costa Norte, edited by Izumi Shimada. Fondo Editorial del Congreso de Perú, Lima. Seoane, F., F.  M. Atoche Castro, K. Chavez Reyes, L. Rodrigo Flores, and G. Rodriquez Romero 2008 Conjunto Arquitectónico No 42: Un Area de Prestación de Servicios en el Sector Norte del Núcleo Urbano Moche. In Informe Técnico 2008, Proyecto Arqueológico Huaca de la Luna, edited by S. Uceda y R. Morales, pp. 253– 307. Universidad Nacional de Trujillo, Trujillo. Shady, Ruth 2009 Caral-Supe y su Entorno Natural y Social en los Origínes de la Civilización. In Andean Civilization: A Tribute to Michael Moseley, edited by Joyce Marcus and Patrick Ryan Williams, pp. 99–120. UCLA Cotsen Institute of Archaeology Press, Los Angeles. Sharon, Douglas 2001 Speculation on Moche Mountain Scenes. In Mortuary Practices and Ritual

4 3 4   Reference List Associations: Shamanic Elements in Prehistoric Contexts in South America, edited by John E. Staller and Elizabeth J. Currie, pp. 109–116. BAR International Series 982. Archaeopress, Oxford, UK. Shay, Talia 1985 Differentiated Treatment of Deviancy at Death as Revealed in Anthropological and Archaeological Material. Journal of Anthropological Archaeology 4:221–241. Shimada, Izumi 1982 Horizontal Archipelago and Coast-Highland Interactions in North Peru: Archaeological Models. In Hombre y su Ambiente en los Andes Centrales, edited by Luis Millones and Hiroyasu Tomoeda, pp. 185–257. Senri Ethnological Series No. 10. National Museum of Ethnology, Suita, Japan. 1986 Batán Grande and Cosmological Unity in the Andes. In Andean Archaeology: Papers in Honor of Clifford Evans, edited by Ramiro Matos, Solveig A. Turpin, and Herbert Eling, pp. 163–188. Monograph 27 of the Institute of Archaeology. University of California Press, Los Angeles. 1990 Cultural Continuities and Discontinuities on the Northern North Coast of Peru, Middle-Late Horizons. In The Northern Dynasties: Kingship and Statecraft in Chimor, edited by Michael E. Moseley and Alana CordyCollins, pp. 297–392. Dumbarton Oaks, Washington, DC. 1994a Pampa Grande and the Mochica Culture. University of Texas Press, Austin. 1994b Los Modelos de la Organizacion Sociopolitical de la Cultura Moche. In Moche: Propuestas y Perspectivas, edited by Santiago Uceda and Elías Mujica, pp. 359–387. Travaux de l’Institut Français d’Etudes Andines, Lima. 1994c Prehispanic Metallurgy and Mining in the Andes: Recent Advances and Future Tasks. In The Quest for Mineral Wealth: Aboriginal and Colonial Mining and Metallurgy in Spanish America, edited by Alan K. Craig and Robert West, pp. 33–73. Louisiana State University Press, Baton Rouge. 1995 Cultura Sicán: Dios, Riqueza y Poder en la Costa Norte del Peru. Edubanco, Lima. 1999 Evolution of Andean Diversity (500 BCE–CE 600). In The Cambridge History of the Native Peoples of the Americas Vol. III, Part 1: South America, edited by Frank Salomon and Stuart B. Schwartz, pp. 350–517. Cambridge University Press, Cambridge, UK. 2000 The Late Prehispanic Coastal States. In The Inca World: The Development of Pre-Columbian Peru, AD 1000–1534, edited by Laura Laurencich Minelli, pp. 49–110. University of Oklahoma Press, Norman. 2009 Who Were the Sicán? Their Development, Characteristics, and Legacies (in Japanese). In Precursor of the Inka Empire: The Golden Capital of Sicán, edited by Izumi Shimada, Ken-ichi Shinoda, and Masahiro Ono, pp. 25–61. Tokyo Broadcasting System, Tokyo. 2010 Moche Sociopolitical Organization: Rethinking the Data, Approaches, and Models. In New Perspectives on Moche Sociopolitical Organization, edited by Jeffery Quilter and Luis Jaime Castillo, pp. 70–82. Dumbarton Oaks, Washington, DC. 2014a Detrás de la Máscara de Oro: La Cultura Sicán. In Cultura Sicán: Esplendor

Reference List  4 3 5 Preinca de la Costa Norte, edited by Izumi Shimada. Fondo Editorial del Congreso de Perú, Lima. 2014b Arte, Religión, y Cosmología Sicán Medio: Nuevos Enfoques. In Cultura Sicán: Esplendor Preinca de la Costa Norte, edited by Izumi Shimada. Fondo Editorial del Congreso de Perú, Lima. Shimada, Izumi (editor) 2014c Cultura Sicán: Esplendor Preinca de la Costa Norte. Fondo Editorial del Congreso de Perú, Lima. Shimada, Izumi, and Raffael Cavallaro 1986 Monumental Adobe Architecture of the Late Prehispanic Northern North Coast of Peru. Journal de la Société des Américanistes LXXI:41–78. Shimada, Izumi, and Alan K. Craig 2013 The Style, Technology, and Organization of Sicán Mining and Metallurgy, Northern Peru: Insights from Holistic Study. Special issue on the “Prehispanic Mining in the Americas.” Chungará: Journal of Chilean Anthropology 45:3–31.​ Shimada, Izumi, and Carlos G. Elera 2007 Informe de la Temporada del Año 2006 del Proyecto Arqueológico Sicán en Huaca Loro en el Sitio de Sicán, Valle Medio de la Leche, Provincia de Ferreñafe, Departamento de Lambayeque. Report on file with the National Institute of Culture, Lima. Shimada, Izumi, and James L. Fitzsimmons 2015 Living with the Dead in the Andes: Introduction. In Living with the Dead in the Andes, edited by Izumi Shimada and James L. Fitzsimmons, pp. 3–49. University of Arizona Press, Tucson. Shimada, Izumi, Adon Gordus, and Jo Ann Griffin 2000 Technology, Iconography, and Significance of Metals: A Multi-Dimensional Analysis of Middle Sicán Objects. In Pre-Columbian Gold: Technology, Iconography, and Style, edited by Colin McEwan, pp. 28–61. British Museum Press, London. Shimada, Izumi, Haagen D. Klaus, Rafael Segura, and Go Matsumoto 2015 Living with the Dead: Conception and Treatment of the Dead on the Central and North Coast of Peru. In Living with the Dead in the Andes, edited by Izumi Shimada and James Fitzsimmons, pp. 101–172. University of Arizona Press, Tucson. Shimada, Izumi, and Adriana Maguiña 1994 Nueva Vision Sobre la Cultura Gallinazo y su Relación con la Cultura Moche. In Moche: Propuestas y Perspectivas, edited by Santiago Uceda and Elías Mujica, pp. 31–58. Travaux de l’Institut Français d’Etudes Andines, Lima. Shimada, Izumi, and Go Matsumoto 2011 Fire, Water, Huaca, and Offerings: Rituals of Regeneration and Ancestor Veneration in the Sicán Culture. Paper presented at the 76th Annual Meeting of the Society for American Archaeology, Sacramento, California. Shimada, Izumi, Crystal Barker Schaaf, Lonnie G. Thompson, and Ellen Mosley-Thompson 1991 Cultural Impacts of Severe Droughts in the Prehistoric Andes:

4 3 6   Reference List Application of a 1,500-Year Ice Core Precipitation Record. World Archaeology 22:247–270. Shimada, Izumi, Rafael Segura Llanos, María Rostworowski de Diez Canseco, and Hirokatsu Watanabe 2005a Una Nueva Evaluación de la Plaza de los Perigrinos de Pachacamac: Aportes de la Primera Campaña 2003 del Proyecto Arqueológico Pachacamac. Bulletin de l’Institut Français d’Études Andines 33:507–538. Shimada, Izumi, Ken-ichi Shinoda, Walter Alva, Steve Bourget, Claude Chapdelaine, and Santiago Uceda 2008 The Moche People: Genetic Perspective on Their Sociopolitical Composition and Organization. In The Art and Archaeology of the Moche: An Ancient Andean Society of the Peruvian North Coast, edited by Steve Bourget and Kimberly L. Jones, pp. 179–193. University of Texas Press, Austin. Shimada, Izumi, Ken-ichi Shinoda, Steve Bourget, Walter Alva, and Santiago Uceda 2005b mtDNA Analysis of Mochica and Sicán Populations of Pre-Hispanic Peru. In Biomolecular Archaeology: Genetic Approaches to the Past, edited by David M. Reed, pp. 61–92. Occasional Paper No. 32, Center for Archaeological Investigations. Southern Illinois University Press, Carbondale. Shimada, Izumi, Ken-ichi Shinoda, Julie Farnum, Robert Corruccini, and Hirokatsu Watanabe 2004 An Integrated Analysis of Pre-Hispanic Mortuary Practices: A Middle Sicán Case Study. Current Anthropology 45:369–402. Shimada, Izumi, and Rafael Vega-Centeno 2011 Peruvian Archaeology: Its Growth, Characteristics, Practice, and Challenge. In Comparative Archaeologies: A Sociological View of the Science of the Past, edited by Ludomir R. Lozny, pp. 569–612. Springer, New York. Shimada, Izumi, and Ursel Wagner 2007 Craft Production on the Pre-Hispanic North Coast of Peru: A Holistic Approach and Its Results. In Archaeology as Anthropology: Theoretical and Methodological Approaches, edited by James Skibo, Michael Grave, and Meriam Stark, pp. 163–197. University of Arizona Press, Tucson. Shimada, Melody, and Izumi Shimada 1981 Explotación y Manejo de los Recursos Naturales en Pampa Grande, Sitio Moche V: Significado del Análisis Orgánico. Revista del Museo Nacional (Lima) 45:19–73. 1985 Prehistoric Llama Breeding and Herding on the North Coast of Peru. American Antiquity 50:3–26. Shipman, Pat 1981 Applications of Scanning Electron Microscopy to Taphonomic Problems. Annals of the New York Academy of Sciences 376:357–385. Shipman, Pat, and Jennie Rose 1983 Early Hominid Hunting, Butchering, and Carcass-Processing Behaviors: Approaches to the Fossil Record. Journal of Anthropological Archaeology 2:57–98. Shulting, Rick, and Linda Fibiger (editors) 2012 Sticks, Stones, and Broken Bones: Neolithic Violence in European Perspective. Oxford University Press, Oxford, UK.

Reference List  4 37 Sillar, Bill 1994 Playing with God: Cultural Perceptions of Children, Play, and Miniatures in the Andes. Archaeological Review from Cambridge 13:47–64. Simmons, Tal 2002 Taphonomy of a Karstic Cave Execution Site at Hrgar, Bosnia-Herzegovina. In Advances in Forensic Taphonomy: Method, Theory, and Archaeological Perspectives, edited by William D. Haglund and Marcella H. Sorg, pp. 263–276. CRC Press, Boca Raton, Florida. Simpson, Keith 1949 Deaths from Vagal Inhibition. Lancet 1:558–560. Sledzik, Paul 1998 Forensic Taphonomy: Postmortem Decomposition and Decay. In Forensic Osteology: Advances in the Identification of Human Remains, edited by Kathleen J. Reichs, pp. 109–119. Charles C. Thomas, Springfield, Illinois. Slovak, Nicole M., and Ann Paytan 2011 Fisherfolk and Farmers: Carbon and Nitrogen Isotope Evidence from Middle Horizon Ancón, Peru. International Journal of Osteoarchaeology 21:253–267. Smith, Bruce N., and Samuel Epstein 1971 Two Categories of 13C/12C Ratios for Higher Plants. Plant Physiology 47:380–384. Smith, Maria Ostendorf 1997 Osteological Indications of Warfare in the Archaic Period of the Western Tennessee Valley. In Troubled Times: Violence and Warfare in the Past, edited by Debra L. Martin and David W. Frayer, pp. 241–266. Gordon and Breach, Amsterdam. Smith, Michael O. 1993 A Probable Case of Decapitation at the Late Archaic Robinson Site (40SM4), Smith County, Tennessee. Tennessee Archaeologist 18:131–142. Sofaer, Joanna 2006 The Body as Material Culture: A Theoretical Osteoarchaeology. Cambridge University Press, Cambridge, UK. Spence, Michael W., Christine D. White, Fred J. Longstaffe, and Kimberley R. Law 2004 Victims of the Victims: Human Trophies Worn by Sacrificed Soldiers from the Feathered Serpent Pyramid, Teotihuacan. Ancient Mesoamerica 15:1–15. Splitstoser, Jeffrey C. 2009 Weaving the Structure of the Cosmos: Cloth, Agency, and Worldview at Cerrillos, an Early Paracas Site in the Ica Valley, Peru. Unpublished PhD dissertation, The Catholic University of America, Washington, DC. Spradley, M. Katherine, and Richard L. Jantz 2011 Sex Estimation in Forensic Anthropology: Skull versus Postcranial Elements. Journal of Forensic Sciences 56:289–296. Sprenger, Guido 2005 The Way of the Buffaloes: Trade and Sacrifice in Northern Laos. Ethnology 44:291–312.

4 3 8   Reference List Steckel, Richard H. 1995 Stature and Standard of Living. Journal of Economic Literature 33:1903–1940. Stojanowski, Christopher M., and Michael A. Schillaci 2006 Phenotypic Approaches for Understanding Patterns of Intracemetery Biological Variation. Yearbook of Physical Anthropology 49:49–88. Strong, William Duncan, and Clifford Evans 1952 Cultural Stratigraphy in the Viru Valley, Northern Peru: The Formative and Florescent Epochs. Columbia University Press, New York. Stuart-Macadam, Patricia L. 1992 Porotic Hyperostosis: A New Perspective. American Journal of Physical Anthropology 87:39–48. Stuiver, M., and P. J. Reimer 1993 Extended 14C Data Base and Revised Calib 3.0 14C Age Calibration Program. Radiocarbon 35:215–230. Sugiyama, Nawa, Raúl Valdez, Gilberto Pérez, Bernardo Rodríguez, and Fabiola Torres 2013 Animal Management, Preparation, and Sacrifice: Reconstructing Burial 6 at the Moon Pyramid, Teotihuacan, Mexico. Anthropozoologica 48:467–485. Sutter, Richard C., and John W. Verano 2007 Biodistance Analysis of the Moche Sacrificial Victims from Huaca de la Luna Plaza 3C: Matrix Method Test of Their Origins. American Journal of Physical Anthropology 132:193–206. Swenson, Edward R. 2003 Cities of Violence: Sacrifice, Power, and Urbanization in the Andes. Journal of Social Archaeology 3:256–296. 2011 Stagecraft and the Politics of Spectacle in Ancient Peru. Cambridge Archaeological Journal 21:285–315. 2012 Moche Ceremonial Architecture as Thirdspace: The Politics of PlaceMaking in the Ancient Andes. Journal of Social Archaeology 12:3–28. Swenson, Edward R., and John P. Warner 2012 Crucibles of Power: Forging Copper and Forging Subjects at the Moche Ceremonial Center of Huaca Colorada, Peru. Journal of Anthropological Archaeology 31:314–333. Szpak, Paul 2013 Stable Isotope Ecology and Human-Animal Interactions in Northern Peru. Unpublished PhD dissertation, Department of Anthropology, University of Western Ontario, London, Ontario. Szpak, Paul, David Chicoine, Jean-François Millaire, Christine D. White, Rebecca J. Parry, and Fred J. Longstaffe In Press Early Horizon Camelid Management Practices in the Nepeña Valley, North-Central Coast of Peru. Environmental Archaeology. Szpak, Paul, Jean-François Millaire, Christine D. White, George F. Lau, Flannery Surette, and Fred J. Longstaffe 2015 Origins of Prehispanic Camelid Wool Textiles from the North and Central

Reference List  4 3 9 Coasts of Peru Traced by Carbon and Nitrogen Isotopic Analyses. Current Anthropology 56:449–459. Szpak, Paul, Jean-François Millaire, Christine D. White, and Fred J. Longstaffe 2012 Influence of Seabird Guano and Camelid Dung Fertilization on the Nitrogen Isotopic Composition of Field-Grown Maize (Zea mays). Journal of Archaeological Science 39:3721–3740. 2014 Small Scale Camelid Husbandry on the North Coast of Peru (Virú Valley): Insight from Stable Isotope Analysis. Journal of Anthropological Archaeology 36:110–129. Szpak, Paul, Christine D. White, Fred J. Longstaffe, Jean-François Millaire, and Víctor F. Vásquez Sánchez 2013 Carbon and Nitrogen Isotopic Survey of Northern Peruvian Plants: Baselines for Paleodietary and Paleoecological Studies. PLoS One 8:e53763. Talavera, Jorge, Juan M. Rojas, and Enrique García 2001 Modificaciones Culturales en los Restos Óseos de Cantona, Puebla: Un Análisis Bioarqueológico. Instituto Nacional de Antropología e Historia, Mexico City. Tanteleán, Henry 2014 Peruvian Archaeology: A Critical History. Left Coast Press: Walnut Creek, California. Tatlock, Jason R. 2006 How in Ancient Times They Sacrificed People: Human Immolation in the Eastern Mediterranean Basin with Special Emphasis on Israel and the Ancient Near East. Unpublished PhD dissertation, Department of Near Eastern Studies, University of Michigan, Ann Arbor. Tello, Julio C. 1937 El Oro de Batán Grande. El Comercio, 18 de abril. Lima. Tello, Ricardo, José Armas, and Claude Chapdelaine 2003 Prácticas Funerarias Moche en el Complejo Arqueológico Huacas del Sol y de la Luna. In Moche Hacia el Final del Milenio, edited by Santiago Uceda and Elias Mujica, pp. 151–188. Universidad Nacional de Trujillo y Pontificia Universidad Católica del Perú, Lima. Tello, Ricardo, and Tania Delabarde 2008 Las Tumbas del Conjunto Arquitectónico 35 de las Huacas del Sol y de la Luna. In Investigaciones en la Huaca de la Luna 2001, edited by Santiago Uceda, Elías Mujica, and Ricardo Morales, pp. 130–173. Facultad de Ciencias Sociales, Universidad Nacional de Trujillo, Trujillo, Perú. Testart, Alain 2006 Interprétation Symbolique et Interprétation Religieuse en Archéologie: L’exemple Taureau à Çatal Höyük. Paléorient 32(2):23–57. Thornton, Erin Kennedy, Susan D. deFrance, John Krigbaum, and Patrick Ryan Williams 2011 Isotopic Evidence for Middle Horizon to 16th Century Camelid Herding in the Osmore Valley, Peru. International Journal of Osteoarchaeology 21:544–567. Tiesler, Vera 2007 Funerary or Nonfunerary? New References in Identifying Ancient Maya

4 4 0  Reference List Sacrificial and Postsacrificial Behaviors from Human Assemblages. In New Perspectives on Human Sacrifice and Ritual Body Treatments in Ancient Maya Society, edited by Vera Tiesler and Andrea Cucina, pp. 14–44. Springer, New York. Tiesler, Vera, and Andrea Cucina 2006 Procedures in Human Heart Extraction and Ritual Meaning: Taphonomic Assessment of Anthropogenic Marks in Classic Maya Skeletons. Latin American Antiquity 17:493–510. Tiesler, Vera, and Andrea Cucina (editors) 2007 New Perspectives on Human Sacrifice and Ritual Body Treatments in Ancient Maya Society. Springer, New York. Tieszen, Larry L., and Michael Chapman 1995 Carbon and Nitrogen Isotopic Status of the Major Marine and Terrestrial Resources in the Atacama Desert of Northern Chile. In Proceedings of the I World Congress on Mummy Studies 1992, Organismo Autónomo de Museos y Centros, pp. 409–425. Museo Arquelógico y Etnográfico de Tenerife, Santa Cruz de Tenerife, Canary Islands, Spain. Tomka, Steve A. 1992 Vicuñas and Llamas: Parallels in Behavioral Ecology and Implications for the Domestication of Andean Camelids. Human Ecology 20:407–433. Toots, Heinrich 1965 Sequence of Disarticulation in Mammalian Skeletons. Contributions to Geology 4:37–39. Topic, John R., and Theresa Lange Topic 2009 Variation in the Practice of Prehispanic Warfare on the North Coast of Peru. In Warfare in Cultural Context: Practice, Agency, and the Archaeology of Violence, edited by Axel E. Nielsen and Willaim H. Walker, pp. 17–55. University of Arizona Press, Tucson. Torres-Rouff, Christina 2011 Hiding Inequality Beneath Prosperity: Patterns of Cranial Injury in Middle Period San Pedro de Atacama, Northern Chile. American Journal of Physical Anthropology 146:28–37. Toyne, J. Marla 2002 Tales Woven in Their Bones: The Osteological Examination of the Human Skeletal Remains from the Stone Temple at Túcume Perú. Unpublished MA thesis, Department of Anthropology, The University of Western Ontario, London, Ontario. 2006 Analysis of Human Skeletal Remains from the Late Intermediate Period Occupation of Farfán, Perú. Technical report on file with the author. 2008 Offering Their Hearts and Their Heads: A Bioarchaeological Analysis of Ancient Human Sacrifice on the Northern Coast of Peru. Unpublished PhD dissertation, Department of Anthropology, Tulane University, New Orleans. 2011a Investigación y Análysis de los Restos Humanos Excavados en Huaca de Las Balsa, Túcume. In Huaca de la Balsas de Túcume: Arte Mural Lambayeque, edited by Alfredo Narváez and Bernarda Delgado, pp. 195–203. Editoral Súper Gráfica, Lima.

Reference List  4 4 1 2011b Interpretations of Pre-Hispanic Ritual Violence at Túcume, Peru, from Cut Mark Analysis. Latin American Antiquity 22(4):505–524. 2015a Ritual Violence and Human Offerings at the Temple of the Sacred Stone, Túcume, Peru. In Living with the Dead in the Andes, edited by Izumi Shimada and James Fitzsimmons, pp. 173–199. University of Arizona Press, Tucson. 2015b The Body Sacrificed: A Bioarchaeological Analysis of Ritual Violence in Ancient Túcume, Peru. Journal of Religion and Violence 3:137–171. Toyne, J. Marla, Christine D. White, John W. Verano, Santiago Uceda, Jean François Millaire, and Fred J. Longstaffe 2014 Residential Histories of Elites and Sacrificial Victims at Huacas de Moche, Peru, as Reconstructed from Oxygen Isotopes. Journal of Archaeological Science 42:15–28. Trever, Lisa, Jorge Gamboa Velásquez, Ricardo Toribio Rodríguez, and Flannery Surette 2013 A Moche Feathered Shield from the Painted Temples of Pañamarca, Peru. Ñawpa Pacha 16:103–118. Trigger, Bruce G. 1990 The Huron: Farmers of the North. Holt, Rinehart and Winston, Fort Worth. 2003 Understanding Early Civilizations: A Comparative Study. Cambridge University Press, Cambridge, UK. Trimborn, Hermann 1979 El Reino de Lambayeque en el Antiguo Perú. Haus Völker und Kulturen Anthropos-Institut, St. Augustin, Germany. Tsai, Howard I. 2012 An Archaeological Investigation of Ethnicity at Las Varas, Peru. Unpublished PhD dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Tschauner, Hartmut 2001 Socioeconomic and Political Organization in the Late Prehispanic Lambayeque Sphere, Northern North Coast of Peru. Unpublished PhD dissertation, Department of Anthropology, Harvard University, Cambridge, Massachusetts. Tschauner, Hartmut, Marianne Vetters, Jalh Dulanto, Marcelo Saco, and Carlos Wester 1994 Un Taller Alfarero Chimú en el Valle de Lambayeque. In Tecnología y Organización de la Producción de Ceramica Prehispanica en los Andes, edited by Izumi Shimada, pp. 349–381. Pontifica Universidad Católica del Peru, Lima. Tufinio, Moisés 2008 Excavaciones en la Plaza 3C de Huaca de la Luna: Nuevas Evidencias de Sacrificios Humanos. In Investigaciones en la Huaca de la Luna 2001, edited by Santiago Uceda, Elias Mujica, and Ricardo Morales, pp. 53–62. Facultad de Ciensas Sociales, Universidad Nacional de Trujillo, Trujillo. Tufinio, Moisés, M. Orbegosa, R. Vega, and C. Rojas 2009 Excavaciones en la Plataforma III de la Huaca de la Luna. In Informe Tecnico 2008, Projecto Arqueologico Huaca de la Luna, complied by

4 4 2   Reference List Santiago Uceda and Ricardo Morales, pp. 113–196. Universidad Nacional de Trujillo, Trujillo. Tung, Tiffiny A. 2005 A River with Parasites Runs through It: Porotic Lesions as Evidence for Iron Loss and Anemia among Three Prehispanic Populations in the Andes of Peru. American Journal of Physical Anthropology Supplement 40:208. 2008 Dismembering Bodies for Display: A Bioarchaeological Study of Trophy Heads from the Wari site of Conchopata, Peru. American Journal of Physical Anthropology 136:294–308. 2012 Violence, Ritual, and the Wari Empire: A Social Bioarchaeology of Imperialism in the Ancient Andes. University Press of Florida, Gainesville. 2014 Agency: ’Til Death Do Us Part? Inquiring about the Agency of Dead Bodies from the Ancient Andes. Cambridge Archaeological Journal. 24(03):437–452 Tung, Tiffiny A., and Mirza del Castillo 2005 Una Visión De La Salud Comunitaria en el Valle de Majes Durante la Época Wari. Corriente Arqueológica 1:149–172. Tung, Tiffiny A., and Kelly J. Knudson 2010 Childhood Lost: Abductions, Sacrifice, and Trophy Heads of Children in the Wari Empire of the Ancient Andes. Latin American Antiquity 21:44–66. 2011 Identifying Locals, Migrants, and Captives in the Wari Heartland: A Bioarchaeological and Biogeochemical Study of Human Remains from Conchopata, Peru. Journal of Anthropological Archaeology 30:247–261. Turner, Bethany L., Haagen D. Klaus, Sarah V. Livengood, Leslie E. Brown, Fausto Saldaña, and Carlos Wester 2013 The Variable Roads to Sacrifice: Isotopic Investigations of Human Remains from Chotuna-Huaca de Los Sacrificios, Lambayeque, Peru. American Journal of Physical Anthropology 151:22–37. Turner, Christy G. II, Christian R. Nichol, and G. Richard Scott 1991 Scoring Procedures for Key Morphological Traits of the Permanent Dentition: The Arizona State University Dental Anthropology System. In Advances in Dental Anthropology, edited by Marc A. Kelley and Clark Spencer Larsen, pp. 13–31. Wiley-Liss, New York. Turner, Christy G. II, and Jacqueline A. Turner 1999 Man Corn: Cannibalism and Violence in the Prehistoric American Southwest. University of Utah Press, Salt Lake City. Turner, Terence 2011 The Body Beyond the Body: Social, Material, and Spiritual Dimensions of Bodiliness. In A Companion to the Anthropology of the Body and Embodiment, edited by Frances E. Mascia-Lee, pp. 102–118. Wiley-Blackwell, New York. Turner, Victor 1977 Sacrifice as Quintessential Process: Prophylaxis or Abandonment? History of Religion 16:189–215. Turvey, Brent 1996 A Guide to the Physical Analysis of Ligature Patterns in Homicide

Reference List  4 4 3 Investigations. Available at: http://www.corpus-delicti.com/ligature .html. Accessed January 1, 2013. Tylor, E. B. 1874 Primitive Culture: Researches in the Development of Mythology, Philosophy, Religion, Language, Arts, and Custom. Holt, New York. 1889 Primitive Culture 2. Holt, New York. Tyson, Rose A., and Alana Cordy-Collins 1998 Taphonomy of a Skull Cache from the North Coast of Peru. Paper presented at the 25th Annual Meeting of the Paleopathology Association, Salt Lake City. Ubbelohde-Doering, Henrich 1952 The Art of Ancient Peru. Praeger, New York. Ubelaker, Douglas H. 1974 Reconstruction of Demographic Profiles from Ossuary Skeletal Samples: A Case Study from the Tidewater Potomac. Smithsonian Institution Press, Washington, DC. 1989 Human Skeletal Remains: Excavation, Analysis, Interpretation. 2nd ed. Aldine, Chicago. 1992 Hyoid Fracture and Strangulation. Journal of Forensic Science 37:1216–1222. 1999 Human Skeletal Remains: Excavation, Analysis, Interpretation. 3rd ed. Taraxcum, Washington, DC. 2002 Approaches to the Study of Commingling in Human Skeletal Biology. In Advances in Forensic Taphonomy: Method, Theory, and Archaeological Perspectives, edited by William D. Haglund and Marcella H. Sorg, pp. 331–351. CRC Press, Boca Raton, Florida. 2008 Methodology in Commingling Analysis: An Historical Overview. In Recovery, Analysis, and Identification of Commingled Human Remains, edited by Bradley Adams and John Byrd, pp. 1–7. Humana Press, New York. Uceda, Santiago 1997 El Poder y la Muerte en la Sociedad Moche. In Investigaciones en la Huaca de la Luna 1995, edited by Santiago Uceda, Elías Mujica, and Ricardo Morales, pp. 177–188. Facultad de Ciencias Sociales, Universidad Nacional de La Libertad, Trujillo, Perú. 1999 Esculturas en Miniatura y una Maqueta en Madera: El Culto a los Muertos y a los Ancestros en la Época Chimú. Beiträge zur Allgemeinen und Vergleichenden Archäologie 19:259–311. 2001 Investigations at Huaca de la Luna, Moche Valley: An Example of Moche Religious Architecture. In Moche Art and Archaeology in Ancient Peru, edited by Joanne Pillsbury, pp. 47–67. National Gallery of Art, Washington, DC. 2010 Theocracy and Secularism: Relationships between the Temple and Urban Nucleus and Political Change at the Huacas de Moche. In New Perspectives on Moche Political Organization, edited by Jeffrey Quilter and Luis Jaime Castillo, pp. 132–158. Dumbarton Oaks, Washington, DC. Uceda, Santiago, and José Armas 1997 Los Talleres Alfareros en el Centro Urbano Moche. In Investigaciones en

4 4 4   Reference List la Huaca de la Luna 1995, edited by Santiago Uceda, Elías Mujica, and Ricardo Morales, pp. 93–104. Facultad de Ciencias Sociales, Universidad Nacional de La Libertad, Trujillo. Uceda, Santiago, and Elías Mujica (editors) 1994 Moche: Propuestas y Perspectivas. Travaux de l’Institut Français d’ Etudes Andines, Lima. 2003 Moche: Hacía el Final del Milenio. Pontificia Universidad Católica del Peru, Lima. Uceda, Santiago, and Moisés Tufinio 2003 El Complejo Arquitectónico Religioso Moche de Huaca de la Luna: Una Aproximación a su Dinámica Ocupacional. In Moche: Hacia el Final del Milenio, edited by Santiago Uceda and Elías Mujica, pp. 179–228. Vol. II. Universidad Nacional de Trujillo, Pontificia Universidad Católica del Perú, Lima. Uceda, Santiago, Moisés Tufinio, and Elías Mujica 2011 El Templo Nuevo de Huaca de La Luna, Primera Parte: Evidencias Recientes Sobre El Moche Tardío. Arkinka, Revista de Arquitectura, Diseño, y Construcción 15:86–98. Uhl, Natalie M., and Stephen P. Nawrocki 2010 Multifactorial Estimation of Age at Death from the Human Skeleton. In Age Estimation of the Human Skeleton, edited by Krista Latham and Michael Finnegan, pp. 243–261. Charles C. Thomas, Springfield, Illinois. Uhle, Max 2014 [1913] Las Ruinas de Moche. Translation by Peter Kaulike. Fondo Editorial Pontificia Universidad Católica del Perú, Lima. 1991 [1903] Pachacamac (A Reprint of the 1903 Edition and Pachacamac Archaeology: Retrospect and Prospect/An Introduction by Izumi Shimada). University Museum of Archaeology and Anthropology, University of Pennsylvania, Philadelphia. Valdez, Lidio M. 2009 Walled Settlements, Buffer Zones, and Human Decapitation in the Acarí Valley, Peru. Journal of Anthropological Research 65:389–416. Valeri, Valerio 1985 Kingship and Sacrifice: Ritual and Society in Ancient Hawaii. Translation by Paula Wissing. University of Chicago Press, Chicago. 1994 Wild Victims: Hunting as Sacrifice and Sacrifice as Hunting in Huaulu. History of Religions 34:101–131. Van Vleet, Krista E. 2002 The Intimacies of Power: Rethinking Violence and Affinity in the Bolivian Andes. American Ethnologist 29:567–601. Verano, John W. 1986a A Mass Burial of Mutilated Individuals at Pacatnamú. In The Pacatnamú Papers, Volume 1, edited by Christopher B. Donnan and Guillermo A. Cock, pp. 117–138. Fowler Museum of Cultural History, University of California, Los Angeles. 1986b H1M1: A Late Intermediate Period Mortuary Structure at Pacatnamú. In The Pacatnamú Papers, Volume 1, edited by Christopher B. Donnan and

Reference List  4 4 5

1992 1995

1997a

1997b 1997c 2000 2001a 2001b 2001c 2003a

2003b 2005

2007

2008a

Guillermo A. Cock, pp. 85–94. Fowler Museum of Cultural History, University of California, Los Angeles. Prehistoric Disease and Demography in the Andes. In Disease and Demography in the Americas, edited by John W. Verano and Douglas H. Ubelaker, pp. 15–24. Smithsonian Institution Press, Washington, DC. Where Do They Rest? The Treatment of Human Offerings and Trophies in Ancient Peru. In Tombs for the Living: Andean Mortuary Practices, edited by Tom D. Dillehay, pp. 189–227. Dumbarton Oaks, Washington, DC. Physical Characteristics and Skeletal Biology of the Moche Population at Pacatnamú. In The Pacatnamú Papers, Volume 2: The Moche Occupation, edited by Christopher B. Donnan and Guillermo A. Cock, pp. 189–214. Fowler Museum of Cultural History, University of California, Los Angeles. Human Skeletal Remains from Tomb 1, Sipán (Lambayeque River Valley, Perú) and their Social Implications. Antiquity 71:670–683. Paleopathology of Andean South America. Journal of World Prehistory 11:237–268. Paleopathological Analysis of Sacrificial Victims at the Pyramid of the Moon, Moche River Valley, Northern Peru. Chungara: Revista de Antropología Chilen0 32:61–70. The Physical Evidence of Human Sacrifice in Ancient Peru. In Ritual Sacrifice in Ancient Peru, edited by Elizabeth Benson and Anita Cook, pp. 165–203. University of Texas Press, Austin. War and Death in the Moche World: Osteological Evidence and Visual Discourse. In Moche Art and Archaeology in Ancient Peru, edited by Joanne Pillsbury, pp. 111–125. National Gallery of Art, Washington, DC. Paleopathological Analysis of Sacrificial Victims at the Pyramid of the Moon, Moche River Valley, Northern Peru. Chungara: Revista de Antropología Chileno 32:61–70. Human Skeletal Remains from Machu Picchu: A Reexamination of the Yale Peabody Museum’s Collections. In The 1912 Yale Peruvian Scientific Expedition Collections from Machu Picchu: Human and Animal Remains, edited by Richard L. Burger and Lucy C. Salazar, pp. 65–171. Yale University Publications in Anthropology, Number 85. New Haven. Mummified Trophy Heads from Peru: Diagnostic Features and Medicolegal Significance. Journal of Forensic Sciences 48:525–530. Human Sacrifice and Postmortem Modification at the Pyramid of the Moon, Moche Valley, Peru. In Interacting with the Dead: Perspectives on Mortuary Archaeology for the New Millennium, edited by Gordon F. M. Rakita, Jane E. Buikstra, Lane Anderson Beck, and Sloan R. Williams, pp. 277–289. University Press of Florida, Gainesville. Conflict and Conquest in Prehispanic Andean South America: Archaeological and Osteological Evidence. In Latin American Indigenous Warfare and Ritual Violence, edited by Richard Chacon and Rubén Mendoza, pp. 105–115. University of Arizona Press, Tucson. Communality and Diversity in Moche Human Sacrifice. In The Art and Archaeology of the Moche: An Ancient Andean Society of the Peruvian North

4 4 6   Reference List Coast, edited by Steve Bourget and Kimberly L. Jones, pp. 195–213. University of Texas Press, Austin. 2008b Trophy Head-Taking and Human Sacrifice in Andean South America. In Handbook of South American Archaeology, edited by Helaine Silverman and William H. Isbell, pp. 1047–1062. Springer, New York. Verano, John W., Laurel Anderson, and Guido P. Lombardi n.d. Analísis Osteológico de los Restos Humanos Moche Hallados en la Huaca Cao Viejo por el Proyecto Arqueológico Complejo El Brujo. Investigaciones en la Huaca Cao Viejo, Valle de Chicama, Perú, edited by Régulo Franco, Cesar Gálvez, and Segundo Vásquez. Fundación Augusto N. Wiese, Lima. Verano, John W., and Heather C. Backo In press Sacrificios Humanos en la Plataforma III de Huaca de la Luna. In Investigaciones en la Huaca de la Luna, edited by Santiago Uceda, Elías Mujica, and Ricardo Morales. Facultad de Ciencias Sociales, Universidad Nacional de la Libertad, Trujillo, Perú. Verano, John W., and Michael J. DeNiro 1993 Locals or Foreigners? Morphological, Biometric, and Isotopic Approaches to the Question of Group Affinity in Human Skeletal Remains Recovered from Unusual Archaeological Contexts. In Investigations of Ancient Human Tissue: Chemical Analysis in Anthropology, edited by Mary K. Sandford, pp. 361–386. Gordon and Breach, Langhorne, Pennsylvania. Verano, John W., and Guido P. Lombardi 1999 Apéndice 3: Análisis del Material Óseo: Tumbas de Cámera Moche en la Plataforma Superior de la Huaca Cao Viejo, Complejo El Brujo. Boletín del Programma Arqueológico El Brujo 1:48–51. Verano, John W., and Jack Rossen 2011 Human Remains. In From Foraging to Farming in the Andes: New Perspectives on Food Production and Social Organization, edited by Tom D. Dillehay, pp. 163–175. Cambridge University Press, Cambridge, UK. Verano, John W. and J. Marla Toyne 2011 Estudio bioantropológico de los Restos Humanos del Sector II, Punta Lobos, Valle de Harmey. In Arqueología de la Costa de Ancash, edited by Miłosz Giersz and Iván Ghezzi, pp. 449–474. ANDES Boletin del Centro de Estudios Precolombinos de la Universidad de Varsovia. Verano, John W., Moisés Tufinio, and Mellisa Lund Valle 2007 Esqueletos Humanos de la Plaza 3C de Huaca de la Luna. In Investigaciones en la Huaca de la Luna 2001, edited by Santiago Uceda, Elais Mujica, and Ricardo Morales, pp. 225–254. Facultad de Ciencias Sociales, Universidad Nacional de Trujillo, Trujillo, Perú. Verano, John W., Santiago Uceda, Claude Chapdelaine, Ricardo Tello, Maria Isabel Paredes, and Victor Pimentel 1999 Modified Human Skulls from the Urban Sector of the Pyramids of Moche, Northern Peru. Latin American Antiquity 10:59–70. Verhoeven, Marc 2002 Ritual and Ideology in the Pre-Pottery Neolithic B of the Levant and Southeast Anatolia. Cambridge Archaeological Journal 12(2):233–258.

Reference List  4 47 Virginia, Ross A., and C. C. Delwiche 1982 Natural 15N Abundance of Presumed N2–fixing and Non-N2–fixing Plants from Selected Ecosystems. Oecologia 54(3):317–325. Voigt, Mary M. 2012 Human and Animal Sacrifice at Galatian Gordion: The Uses of Ritual in a Multiethnic Community. In Sacred Killing: The Archaeology of Sacrifice in the Ancient Near East, edited by Anne Porter and Glenn M. Schwartz, pp. 237–290. Eisenbrauns, Winona Lake, Indiana. Voorhies, Michael 1969 Taphonomy and Population Dynamics of an Early Pliocene Vertebrate Fauna, Knox County, Nebraska. Contributions to Geology Special Paper 2:1–69. University of Wyoming Press, Laramie. Vreeland, James M. 1998 Mummies of Peru. In Mummies, Disease, and Ancient Cultures, 2nd ed., edited by Aidan Cockburn, Eve Cockburn, and Theodore A. Reyman, pp. 154–189. Cambridge University Press, Cambridge, UK. Walde, Héctor 1998 Informe Final: Proyecto Arqueológico en Punta Lobos Puerto de Huarmey: Compañía Minera Antamina S.A. Technical report on file with the Instituto Nacional de Cultura, Lima, Peru. Waldron, Tony 1994 Counting the Dead: The Epidemiology of Skeletal Populations. Wiley-Liss, Chichester, UK. Walker, Phillip L. 2001 A Bioarchaeological Perspective on the History of Violence. Annual Review of Anthropology 30:573–596. Walker, Phillip L., Rhonda R. Bathurst, Rebecca Richman, Thor Gjerdrum, and Valerie A. Andrushko 2009 The Causes of Porotic Hyperostosis and Cribra Orbitalia: A Reappraisal of the Iron-Deficiency-Anemia Hypothesis. American Journal of Physical Anthropology 139:109–125. Walker, Phillip L., and Jeffrey C. Long 1977 An Experimental Study of the Morphological Characteristics of Tool Marks. American Antiquity 42:605–616. Walker, William H. 1995 Ceremonial Trash? In Expanding Archaeology, edited by James M. Skibo, William H. Walker, and Axel E. Nielsen, pp. 67–79. University of Utah Press, Salt Lake City. Walker, William H., and Lisa J. Lucero 2000 The Depositional History of Ritual and Power. In Agency in Archaeology, edited by Marcia Anne Dobres and John E. Robb, pp. 130–147. Routledge, London. Weber, Jill A. 2012 Restoring Order: Death, Display, and Authority. In Sacred Killing: The Archaeology of Sacrifice in the Ancient Near East, edited by Anne Porter and Glenn M. Schwartz, pp. 159–190. Eisenbrauns, Winona Lake, Indiana.

4 4 8   Reference List Webster, Steven 1972 The Social Organization of a Native Andean Community. Unpublished PhD dissertation, Department of Anthropology, University of Washington, Seattle. Weismantel, Mary 2015 Many Heads Are Better than One: Mortuary Practice and Ceramic Art in Moche Society. In Living with the Dead in the Andes, edited by Izumi Shimada and James L. Fitzsimmons, pp. 76–100. University of Arizona Press, Tucson. Weiss-Krejci, Estella 2003 Victims of Human Sacrifice in Multiple Tombs of the Ancient Maya: A Critical Review. In Antropología de la Eternidad: La Muerte en la Cultura Maya, edited by Andrés Ciudad, Mario Humberto Ruz, and María Josefa Ponce de León, pp. 355–381. Sociedad Española de Estudios Mayas, Madrid. Wester, Carlos 1996 Proyecto de Rescate Arqueológico de Úcupe-Pueblo. Technical report on file at the Museo Nacional de Arqueología y Etnografía Hans Henrich Brüning, Lambayeque, Peru. 2012 Sacerdotisa Lambayeque de Chornancap: Misterio e Historia. Ministerio de Cultura, Lima. Wester, Carlos (editor) 2010 Chotuna-Chornancap: Templos, Rituales, y Ancestros Lambayeque. Editoral Súper Gráfica, Lima. Wester, Carlos, Manuel Curo, and Denis Echeverria 2010b Huaca Chornancap. In Chotuna-Chornancap: Templos, Rituales, y Ancestros Lambayeque, edited by Carlos Wester, pp. 151–183. Editoral Súper Gráfica, Lima. Wester, Carlos, Fausto Saldaña, Samuel Castillo, and Haagen D. Klaus 2010a Huaca de los Sacrificios. In Chotuna-Chornancap: Templos, Rituales, y Ancestros Lambayeque, edited by Carlos Wester, pp. 58–108. Editoral Súper Gráfica, Lima. Wheeler, Jane C. 1982 Aging Llamas and Alpacas by Their Teeth. Llama World 1(2):12–17. Wheeler, Jane C., A. J. F. Russel, and Hilary Redden 1995 Llamas and Alpacas: Pre-Conquest Breeds and Post-Conquest Hybrids. Journal of Archaeological Science 22:833–840. White, Christine D., Andrew J. Nelson, Fred J. Longstaffe, Gisela Grupe, and Anna Jung 2009 Landscape Bioarchaeology at Pacatnamu, Peru: Inferring Mobility From δ13C and δ15N Values of Hair. Journal of Archaeological Science 36:1527–1537. White, Theodore E. 1953 A Method of Calculating the Dietary Percentage of Various Food Animals Utilized by Aboriginal Peoples. American Antiquity 4:396–398. White, Tim D. 1992 Prehistoric Cannibalism at Mancos 5MTUMR-2346. Princeton University Press, Princeton, New Jersey.

Reference List  4 4 9 White, Tim D., Michael T. Black, and Pieter A. Folkens 2012 Human Osteology, 3rd ed. Elsevier, Amsterdam. Wiget, Andrew, and Olga Balalaeva 2001 Khanty Communal Reindeer Sacrifice: Belief, Subsistence, and Cultural Persistence in Contemporary Siberia. Arctic Anthropology 38:82–99. Wild, E. M., K. A. Arlamovsky, R. Golser, W. Kutschera, A. Priller, S. Puchegger, W. Rom, P. Steier, and W. Vycudilik 2000 14C Dating with the Bomb Peak: An Application to Forensic Medicine. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 172(1–4):944–950. Wilke, G. 1933 Bestattung in Bauchlage und verwandte Brauche. In Homenagem a Martins Sarmento: Miscellanea de Estudos em Honra do Investigador Vimaranense, no Centenario do seu Nascimento (1833–1933), edited by Francisco Martins, pp. 449–460. Sociedade Martins Sarmento, Guimaras. Willerslev, Rane 2004 Not Animal, Not Not-Animal: Hunting, Imitation, and Empathetic Knowledge among the Siberian Yukaghirs. Journal of the Royal Anthropological Institute 10:629–652. 2007 Soul Hunters: Hunting, Animism, and Personhood among the Siberian Yukaghirs. University of California Press, Berkeley. Willey, Gordon R. 1953 Prehistoric Settlement Patterns in the Virú Valley, Perú. Smithsonian Institution Press, Washington, DC. Williams, Howard M. R. 2004 Death Warmed Up: The Agency of Bodies and Bones in Early AngloSaxon Cremation Rites. Journal of Material Culture 9:263–291. Williams, Jocelyn S., and M. Anne Katzenberg 2012 Seasonal Fluctuations in Diet and Death During the Late Horizon: A Stable Isotopic Analysis of Hair and Nail from the Central Coast of Peru. Journal of Archaeological Science 39:41–57. Wilson, Andrew S., Timothy Taylor, Maria Constanza Ceruti, Jose Antonio Chavez, Johan Reinhard, Vaughan Grimes, Wolfram Meier-Augenstein, Larry Cartmell, Ben Stern, Michael P. Richards, Michael Worobey, Ian Barnes, and M. Thomas P. Gilbert 2007 Stable Isotope and DNA Evidence for Ritual Sequences in Inca Child Sacrifice. Proceedings of the National Academy of Sciences 104:16456– 16461. Winsborough, Barbara, Izumi Shimada, Lee A. Newsom, John Jones, and Rafael A. Segura 2012 Paleoenvironmental Catastrophes on the Peruvian Coast Revealed in Lagoon Sediment Cores from Pachacamac. Journal of Archaeological Science 39:602–614. Winterhalder, Bruce, Robert Larsen, and R. Brooke Thomas 1974 Dung as an Essential Resource in a Highland Peruvian Community. Human Ecology 2:89–104.

4 5 0  Reference List Wood, James W., George R. Milner, and Henry C. Harpending 1992 The Osteological Paradox: Problems of Inferring Prehistoric Health from Skeletal Samples. Current Anthropology 33:343–370. Wright, Lori E., and Cassady J. Yoder 2003 Recent Progress in Bioarchaeology: Approaches to the Osteological Paradox. Journal of Archaeological Research 11:43–70. Yakir, Dan 2011 The Paper Trail of the δ13C of Atmospheric CO2 since the Industrial Revolution Period. Environmental Research Letters 6:1–4. Yoffee, Norman 2005 Myths of the Archaic State: Evolution of the Earliest Cities, States, and Civilizations. Cambridge University Press, Cambridge, UK. Yoshida, Bonnie 2004 Status and Health amid Changing Social Conditions: Bioarchaeology of a Prehispanic Moche Valley Population. Unpublished PhD dissertation, Department of Anthropology, University of California, Santa Barbara. Yuan, Jing, and Rowan Flad 2005 New Zooarchaeological Evidence for Changes in Shang Dynasty Animal Sacrifice. Journal of Anthropological Archaeology 24:252–270. Zárate, Agustín de 1968 [1556] The Discovery and Conquest of Peru. Translation by John Cohen. Penguin, Baltimore. Zevallos, Jorge 1971 Cerámica de la Cultura “Lambayeque” (Lambayeque 1). Universidad Nacional de Trujillo, Trujillo, Perú. 1989 Introduccion a la Cultura Lambayeque. In Lambayeque, edited by José Antonio de Lavalle, pp. 15–104. Colección Arte y Tesoros del Peru, Banco del Credito, Lima. Zighelboim, Ari 1995 Mountain Scenes of Human Sacrifice in Moche Ceramic Iconography. Journal of the Steward Anthropological Society 23(1–2):153–188. Zuidema, R. Tom 1977–1978 Shafttombs and the Inca Empire. Steward Anthropological Society Journal 9(1–2):133–178. 1990 Dynastic Structures in Andean Cultures. In The Northern Dynasties: Kingship and Statecraft in Chimor, edited by Michael E. Moseley and Alana Cordy-Collins, pp. 489–505. Dumbarton Oaks, Washington DC. 1992 Inca Cosmos in Andean Context. In Andean Cosmologies through Time, edited by R. Dover, K. Seibold, and J. McDowell, pp. 17–45. Indiana University Press, Bloomington.

Index  4 5 1

in d e x

Page numbers followed by f or t indicate material in figures or tables. aclla: weavers, 279; women in service in Inka Sun Temples, 268 acompanante (retainer/companion), 266. See also retainer/companion killing adult stature, 224–225, 231 age and sex of victims: age estimation techniques, 157–158; implications of, 3, 264; retainer/companion killing, 128, 139, 285–289; sacrificed camelids, 330, 335–337. See also individual sites agriculture, 126, 145, 205 Akapana (Tiwanaku), 356 alpaca, 320–322, 341n2. See also camelids Alva, Walter, 17, 91, 246, 249, 250, 288 ancestors, veneration of, 94, 127, 140, 350, 366; gift economy with, 323; propitiating or feeding of, 145, 172 ancient DNA (aDNA), 242, 246, 263, 290, 340 anemia, 138, 161, 183, 224, 228–230 animals: birds, 103; cattle, 176; dogs, 135, 176, 325, 348; domesticated versus wild, 324, 325, 366; felines, 8, 18f, 104, 115, 136, 365; guinea pigs, 325; “ordinary” vs. ritual killings of, 365–366; pigs, 176, 321; ritual bleeding/killing of, 101, 322–324; sacrifice associated with predators, 104; sacrifice of, 322–324; snakes and strangulation, 103–106 (104f, 105f, 106f ), 115, 365; as substitutes for humans, 172, 324. See also camelids antemortem injuries, abnormalities,

160, 225; Huambacho, 111; Pacatnamú, 259, 263; Punta Lobos, 260–261; Pyramids of Moche, 239, 250, 254–255, 263–264; Santa Rita B CA3, 159; Templo de la Piedra Sagrada, 231–234 (232t, 233f, 234f ), 238; tooth loss, 184, 187, 195, 203f archaeothanatology, 372, 377 architecture, 247 Arriaga, Pablo Joseph de, 239 arsenic, 123 art history, 5, 67, 120–121 artistic traditions, 245–246; depicting captives/prisoners, 18f, 244, 248f, 253f, 256f; depicting decapitation, 15–16 (16f, 17f ), 19, 30, 125–126 (126f ), 248f, 205; depicting genital mutilation, 30, 61; depicting ligatures/bindings, 104–106 (104f–106f ), 115; depicting throat-slitting, 21f Aspero, 14 auto-sacrifice/suicide, 3, 15, 289 Benson, Elizabeth P., 5, 100, 103, 125, 364 Betanzos, Juan de, 100, 239, 267–268, 289 bindings. See ligatures/bindings Binford, Lewis, 33–34 bioarchaeology, 66, 96, 221, 243, 280, 372, 375–379; and childhood metabolic stress, 224; of ChotunaChornancap, 179, 183; evidence of warfare/warriors, 37, 225, 239–241, 247–250, 255, 257–258, 261; identifying victims, 179, 212, 269, 280; of retainer/companion killing, 128, 268; of strangulation, 97–98, 99 biocultural models, 21–22, 236–237, 279, 280, 285

4 5 2   Index biodistance analysis, 242, 280–281 (281t), 285, 369 birth/pregnancy, 146, 206, 273, 276, 277f, 284, 288 Bloch, Maurice, 116–117 blood sacrifice: adult female victims of, 204, 205; among Moche, 101; among Muchik, 145–147; versus captive execution, 367; collected in goblets, 31, 58, 101, 126; and Copulation Ceremony, 105; versus funerary/invisible sacrifice, 139–140, 287; hair and, 208; laceration cut marks, 55–56; power of, 342–343, 378; versus strangulation, 98, 99, 102–105 (104f ), 115; in Warrior Narrative, 30. See also Sacrifice Ceremony/Presentation Theme; throat-slitting body concepts, 378–379 body parts, 292; versus bodies, 142; of camelids, 145; ceramic representations of, 31; collection of, 88; display of, 15, 61, 90–91, 101, 117, 142, 263; manipulation of, 95; missing, 61; as offerings, 91, 100, 116, 339; recycling of, 373; re-interment of, 152, 160–161, 163–167, 169, 176, 374; scattering of, 91; as trophies, 263, 373; vital essences in, 376, 378–379. See also disarticulation/ dismemberment body position, 69f, 70, 71f, 165–167 (166f ), 250f–251f, 293 bone collagen isotopic data: at Huaca de Los Sacrificíos, 184–185 (184f ), 195–196; at Huancoco (llamas), 330, 331–333 (332t, 333f ), 335, 336–337; at Pacatnamú, 259; at Santa Rita B, 167 bones: and adult stature, 224–225; as grave goods, 91, 96; os coxae relocation, 302, 306; pit pattern on, 92–94 (93f ); skeletal pathology and trauma, 158–160. See also cut mark analysis boot-shaped tombs, 274, 295, 296–298

bronze, 11, 123–124, 127, 217 Brüning Museum, 179, 182 Buikstra, Jane E., 127 burial(s): camelids buried with children, 20, 153, 161, 192, 200; canchero ritual burials, 351, 352f, 354f; of children, 267, 275–276, 286–287; deviant burials, 116, 140; of fetus, 273 (273f ); Huaca Loro, 123f, 124, 127; Huambacho graves, 109–115 (110f, 112f, 113f, 114f ); “like with like,” 152–153, 169, 170t, 200, 374; mass burial commingling and physical complexity, 65–67; prolonged burial/delayed interment, 67, 88, 94, 129, 201, 271–275, 288, 307; re-interment, 152–153, 165, 169, 170t, 193, 374; with sumptuary goods, 262; without honor, 240. See also “like with like” pattern Burial Theme, 19, 103 burning, 14–15, 23, 310, 311f, 323, 347 CA3 (Conjunto Arquitectónico 3), 154– 155, 156f, 157f, 158f, 160, 169, 174 Cabello de Balboa, Miguel, 12, 180, 182 cache remains, 375 Calancha, Antonio de la, 103, 146, 205 calculated sacrifice, 172–173, 366 camelids: age and sex of sacrificed, 330, 335–337; body parts, 124, 145, 155, 190, 325; buried with children, 20, 153, 161, 192, 200; at Cerro Cerrillos, 133; chicha feedings before sacrifice, 321; coastal husbandry of, 321–322; dietary isotope ratios, 327; at Dos Cabezas, 275; as funerary feasts, 316; geographic origin of, 327, 335–336; herding of, 320–321; at Huaca Loro, 279; at Huancaco, 319–320, 326, 329–337 (329f, 331f, 332t, 333f, 334f ); by Inka, 325; isotopic analysis of bone, hair, 330–337 (332t, 333f, 334f ), 340; manner of death of, 326; at M-U1525, 299; Santa Rita B, 152, 154–155, 158f, 161,

Index  4 5 3 167–168 (168f ), 171–176, 326; selection of for sacrifice, 336–339, 340; as substitutes for humans, 172, 324. See also llamas canchero ritual burials, 351, 352f, 354f, 355f cannibalism, 14 captive executions: archaeological evidence of, 244; different from sacrifice, 367; motives for, 262–264; ritual tribute versus, 367. See also warfare/warriors captives/prisoners: antemortem injuries in, 250–252; depictions of, 18f, 244, 248f, 253f, 256f; Early Intermediate Period/Middle Horizon, 249–261; identifying from bindings, 263; identifying from desecration, exposure of corpses, 262–263; identifying from skeletal remains, 246–247; identifying in art, 245– 246; Late Horizon, 261–262 carnivore scavenging, 84, 86–87, 89f “cathected object,” 90 CE (Chapman Estimator), 73, 77–78 (78t) cemeteries: looted sites, 247, 257; for Sícan elites, 123–125, 127–129, 134, 139; Túcume, 223, 226–236 (226t, 227f–229f, 230t, 232t, 233f–234f, 236f ) Central Coast of Peru, 7f, 14, 288 ceramic jewelry, 306–307 ceramic vessels: burial of, 135; burial under firing area, 131, 132f; depicting burial, 109; depicting captives, 16, 18f, 30, 101, 247–249 (248f ), 253f; depicting combat, 16, 17f, 101; depicting Decapitators, 16, 17f, 19; depicting sacrifice, 15, 31, 368; Gallinazo/Virú, 107, 109, 351–354; as grave goods, 111, 124, 200, 206, 277–278, 296–297; at Huaca de la Luna, 375; intentional destruction of, 15, 136 (136f ); ligatures on, 102– 103; Moche period styles, 107, 109, 293–295 (294f ); Pampa de Burros

workshop, 180; placed to collect rainwater, 143, 144f; sacrifices of, 102–103, 357, 363, 365; stirrup-spout vessels, 20, 295; tinajas and porrones, 135 Cerro Blanco, 19, 35 (35f ), 67–68. See also Huaca de la Luna Cerro Cementerio, 343–344 (344f ) Cerro Cerrillos (Lambayeque valley), 6f, 131–134 (133f ), 145–146, 222, 326, 357 Cerro de Huarpe (Virú Valley), 326 Cerro La Raya, 216. See also Templo de la Piedra Sagrada Cerro Sechín, 6f, 15, 247, 248f chamber tombs, 295, 296–298 Chan Chan, 6f, 11, 20, 179, 262, 279, 357 Chapman Estimator (CE), 73, 77–78 (78t) Chavín de Huantar, 15 chest-opening: for heart removal, 199; at Huaca de la Luna Platform III, 144; at Huaca de la Luna Plaza 3C, 48; Huaca de los Sacrificios, 198–199; Pacatnamú mass grave, 72; Templo de la Piedra Sagrada, 219 (219f ) Chicama, 19, 108, 121, 275, 353. See also El Brujo chicha (maize beer), 135, 267, 295, 296, 321 children: anemia in, 138, 161, 183, 224, 228–230; burials of, 275–276, 286–287; camelids buried with, 20, 153, 161, 192, 200; dedicatory burial at Plaza 3A, 140–141; documenting skeletal health of, 224–226; foundation/construction sacrifices of, 14; guardians in tombs of, 297; immolations, 374; “like with like” burials, 153; live burial in Ecuador, 267; as mallquis/huaca interlocutors, 146; metabolic disruption in, 230–231; as “not quite human,” 146; in pit tombs, 296; sacrificed to moon, 205; sacrifice through time, 170t

4 5 4   Index Chimú culture: conquest of Lambayeque valley, 179–180; depiction of captives, 244; as expansionist empire, 264; Inka defeat of, 12, 180, 261; and Pacatnamú/Punta Lobos mass executions, 261, 263; retainer/companion killing, 279; stirrup-spout vessel, 20, 21f. See also Chotuna-Chornancap; Templo de la Piedra Sagrada (TPS), Túcume China, 176 Chinchasuyu, 180 Chotuna-Chornancap, 6f, 181f, 182–183, 326; age and sex of victims, 183, 185–186, 190, 202f, 204; belief and sacrificial symbolisms, 205–207; and Chimú occupation, 182; Chornancap Norte, 6f, 185–190 (186f–189f ), 278, 285; huacas within, 182; human sacrifices, 183; sequence of ritual killings, 200–203 (201f–203f ); site chronology, 182; victim demography, life histories, and identities, 203–205 chronology, Peruvian referent, 7f Cieza de León, Pedro de, 100, 145, 205, 239, 240, 267 cinnabar pigment, 145, 276 climatic disturbance, 70, 139. See also ENSO (El Niño Southern Oscillation) events Cobo, Bernabe, 13, 145, 207, 239–240, 267, 289, 325 coca leaves, 13, 110 coffins, 271–272 (272f ), 297–298, 300–301 Colle’s fracture, 231, 233 Colonial Period, 170t; accounts of Inka executions, sacrifices, 100, 262, 377; biological stress in, 222; burials during, 170t, 182, 216–217; female attendants called “widows,” 286; looting at Chan Chan, 279; meaning of huaca, 24n1. See also Muchik peoples columnar boxes/sockets, 127, 128f, 139–141

commoner burials, 123f, 124, 145, 285, 295, 296 companion sacrifice. See retainer/ companion killing Conchopata, 356–357 conopa figurine, 191f, 192 construction sacrifices. See foundation/construction sacrifices copper, 123, 217; bells, 17f; bracelet, 306; conopa figurine, 191f, 192; disks, 111, 271; in hands and mouth, 111, 114f, 115; headdress, crown, 271–272; knife, tumi, 31, 302, 305; nose clip, 275; sheets, 115, 300 cosmovisions, 125, 126f, 143, 206, 316 cotton: blindfold, 127; items wrapped in cloth, 111; ligatures, 101, 113–115, 127; pouch, 348; sash, 276; seeds, 200; shrouds, 110, 113, 199, 217; yarns, 111 cradle-boarding, 225, 235 cranial modification, 222, 225–226, 235 (235t), 238 cranial trauma: at El Castillo de Santa, 19, 108; at Huanchaco, 20; Pacatnamú, 72, 85f, 86; Plaza 3A, 38, 68, 90, 101 cribra orbitalia, 224; at Túcume, 228– 230 (228f, 230t), 237 “crisis sacrifices,” 140 crisoles, 299 Cupisnique society, 8, 15, 16f, 24, 104, 245f, 247, 248f curaca/kuraka (lord): children buried with, 267; Chimú lords, 279; competing for power, 173, 180; holding power of life and death, 4; Lord of Úcupe, 273, 288–289; relationship with retainers, 142–143, 267, 285–286; Sicán lords, 123, 129, 139, 271–273 curers punished for failure, 103 cut mark analysis, 32–34, 62–63, 240, 365, 368; on animals, 63, 171, 307, 312; at Cao Viejo, 91; at Cerro Cerrillos, 132–133 (133f ); at El Castillo, 108; at Huaca Chornancap Norte,

Index  4 5 5 188–190 (188f, 189f ), 201–202; at Huaca de la Luna, 31, 38–39; at Huaca de la Luna Plaza 3A, 40t–41t, 48–53 (50f, 52f, 53f ), 54t, 68, 88, 252 (252f ); at Huaca de la Luna Plaza 3C, 39, 40t–43t, 44–48 (44f, 45f, 47f ), 51–53 (51f, 52f ), 54t, 68, 88, 252 (252f ); at Huaca de los Sacrificios, 196f, 197f, 198–199; at Huaca las Ventanas, 129; at Matrix 101, 137; at Pacatnamú, 72, 87–88, 94, 257; at Punto Lobos, 259–260 (260f ); at Santa Rita B, 164 (164f ), 168, 171 cut marks on bone, 55f, 58f; on animal bones, 307, 312; attachment points, 33; calcaneus, 46; carpals, 41t, 43t, 46, 87; cervical vertebrae, 32, 39, 40t, 42t, 44, 48, 49, 51–53, 57, 59, 129, 133, 188, 189f, 197f, 198, 252, 259; clavicle, 39, 40t, 42t, 44, 46, 48, 49, 133, 188, 197f, 198, 259; directionality of, 34, 38–39, 51–54 (51f, 54t), 57 (57f ), 61; femur, 33, 39, 41t, 43t, 46–49, 51, 86, 91; fibula, 39, 41t, 43t, 46, 48, 49, 54, 59; frontal bone, 40t, 42t, 44, 49; humerus, 33, 40t, 42t, 45, 48, 49, 51; ilium, 32, 46; ischium, 46; long bones, generally, 33, 59; lumbar vertebrae, 40t, 42t, 44; malar, 40t, 42t; mandible, 32–33, 40t, 42t, 56; manubrium, 188, 189f, 198; maxilla, 40t, 42t; metacarpals, 41t, 43t, 45, 49, 50; metatarsals, 41t, 43t, 46; nasal bone, 40t, 42t; occipital bone, 40t, 42t, 44, 49, 51; os coxae, 41t, 43t, 46–49, 54; parietal bones, 40t, 42t, 44, 49; patella, 41t, 43t, 48; pelvis, 33, 46; phalanges, 41t, 43t, 45, 46, 48, 49, 50, 54; pubis, 46; radius, 33, 40t, 42t–43t, 45, 46, 49, 87, 257; ribs, 32, 39, 40t, 42t, 44, 48, 49, 54, 72, 133, 164, 168, 188, 190, 198, 259, 260; sacrum, 41t, 43t; scapula, 32, 39, 40t, 42t, 45, 46, 48, 49, 53f; shoulder effects and barbs, 34; skull, 32,

48; sternum, 40t, 42t, 72, 133, 137, 164, 188; tarsals, 41t, 43t, 48; temporal bone, 40t, 42t, 44; thoracic vertebrae, 32, 40t, 42t, 44, 46, 188, 259, 260f; tibia, 41t, 43t, 46, 48, 49; ulna, 41t, 43t, 45, 49, 87, 164 decapitation: of ceramic figures, 136 (136f ); Cerro Cerrillos, 133 (133f ); Chornancap Norte, 188, 206; in Cupisnique art, 104; Decapitator figures/theme, 15–16 (16f ), 17f, 19, 30, 125–126 (126f ), 205; depictions involving severed heads, 125–126 (126f ), 248f; El Castillo, 108; Huaca de la Luna, 44 (44f ), 49, 56, 59, 252; Huaca Las Ventanas, 129; in Middle Sicán, 129; Pacatnamú, 72, 257; partial/semi, 133, 188, 201; signatures of, 32; TPS, 219, 239–241; trophy heads, 252, 262; at Túcume, 20; use of tumi for, 56 decomposition/decay, stages of, 65–66, 88, 91, 135, 199–200, 288, 301–302 dedicatory offerings, 19; of both humans and nonhuman objects, 355–357; at Chan Chan, 223; dismemberment of, 312; at Huaca Santa Clara, 350; humans, 23, 100, 312; needs and means affecting choice of, 357; possible depiction of, 140; range of, 147; at Sicán, 127–129 (128f ), 139–141; at Templo de la Piedra Sagrada, 223. See also foundation/construction sacrifices; propitiatory ritual definitions, 2–5 defleshing: cut marks indicating, 32–33; at Huaca de la Luna, 44–46 (45f ), 49–50, 102; skeletons and, 55–56, 61, 119n2 De la Vega, Garcilaso, 13 delayed interment, 129, 201, 288. See also re-interment of body parts/ secondary interments; reopening of graves

4 5 6   Index dental biodistance data, 280–281 (281t) dental health, 134, 138, 184 (184f ), 186–187, 195, 203f, 204, 286, 230–231 Denver Kero, 125 desecration of corpses, 262, 265 deviant burials, 116, 140 de Xérez, Francisco, 240 diet, determining: Chotuna, 208; dietary isotope ratios, 327; teeth, bones, hair, 184–185 (184f ) directionality of cuts, 34, 38–39, 51–54 (51f, 54t), 57 (57f ), 61 disarticulation/dismemberment, 30–33, 61, 145; agents of alteration of, 66–67; anatomical mapping of, 75; Burial M-U1221, 312; by carnivores, 84; at El Castillo, 108; at Huaca de la Luna Plaza 3A, 50, 82t, 86, 88–94 (89f, 90f, 93f ); at Huaca de la Luna Plaza 3C, 46–48 (47f ), 56; at Pacatnamú, 83t, 86–87, 94–95, 257; rate variability due to climate, 66; in Sacrifice Ceremony, 91–92; at Santa Rita B, 171; and taphonomic winnowing, 84–88; for trophy collection, 94 display of body parts, 15, 19, 61, 90, 117, 142, 263 DNA: ancient DNA (aDNA), 242, 246, 263, 290, 340; mitochondrial DNA (mtDNA), 123–124, 254, 285 DNI (derived number of individuals), 74, 77–79 (78t), 80t Donnan, Christopher: Burial Theme/ Ceremony, 19, 103; ChotunaChornancap excavation, 182; depiction of captives, 246, 249; Dos Cabezas excavation, 274–275; Huanchaco, 20, 153; Moche state religion, 29, 121, 151; Moche “Warrior Narrative,” 17f–18f, 30–31, 101, 104f, 249, 264; religious sacrifice, 20. See also Sacrifice Ceremony/ Presentation Theme

Dos Cabezas, 6f, 97, 153, 274–275, 292 early civilization in Andes, 7–8 Early Horizon, 14, 109–110, 170t, 322 Early Intermediate Period, 169, 170t; camelids in, 152, 168f, 171, 321–323, 329, 334, 337–339; Huambacho, 6f, 109–115 (110f, 112f, 113f, 114f ); “like with like” burials, 153; possible ritual strangulation, 108–109; Virú state system, 343 El Brujo (Huaca Cao Viejo), 6f, 19–20, 30, 91, 170t, 270t, 275 El Castillo de Santa, 6f, 19, 107, 108, 117 elites: aclla, 279; alliances among, 9; ancestor cult among, 10, 127, 134, 147; and animal sacrifice, 316; in artwork, 102; burial according to status, 107; burial traditions of, 270; and cranial modification, 225; endogamy among, 286; Gallinazo, 107–108; health, identity, and social status, 236–238; health of, 222, 237; and Huaca de la Luna, 9, 35, 90; intact, unmutilated bodies, 27; metalwork in burials of, 125; Moche funerals/burials, 30, 37, 60, 111, 292, 295, 299, 329; and “Moche state religion,” 151; power through ritual, 4, 176, 238; prestige sacrifices, 173; and priestess cult, 299; residential structures for, 345, 347; ritual combat among, 254; scholarship on burials, 116; Sícan funerals/cemeteries for, 123–125, 127–129, 134, 139; Sícan mass sacrifice, 138; Sícan tombs of, 271, 274, 276–277, 279; strategies/ goals of, 147; unnatural death of, 293. See also retainer/companion killing El Paraíso, 356 empire building, 11–12, 20, 178–179, 216, 264

Index  4 5 7 enamel hypoplasias, 183, 184f, 186, 224, 237 ENSO (El Niño Southern Oscillation) events, 101; 600 BC event, 8; AD 550 event, 10, 35, 60, 121; AD 1050– 1100 event, 11, 134, 135f, 137–138, 142, 172, 254 entombment: biodistance analysis, 280–284 (281t); boot-shaped tombs, 274, 295, 296–298; chamber tombs, 295, 296–298; “guardians” of tombs, 271, 274, 288, 297, 298; Huaca Lucía “temple entombment,” 15; live entombment, 137; pit tombs, 295, 296; Sícan elites, 271, 274, 276–277, 279; “tomb choreography,” 268; tomb of Lord of Úcupe, 273 (273f ), 288 epigraphy, 246, 378 ethnohistory, 12–14, 316; Chimú defeat of Inka, 261; of fly larval infestation, 206; of high-altitude herders, 320; of Inka animal sacrifice, 321–322, 325; of Inka human sacrifice, 216, 239–242, 262; of Jequetepeque valley, 295; of retainer burials, 267–269, 286, 287, 289; of ritualized violence, 372, 374–375; of strangulation, 100, 287 experimental studies, 33, 92–94 (93f ), 135 Farfán, 6f, 230t, 235t, 236–238 feet: absence of from natural causes, 305; binding of, 102, 113, 115, 263; mutilation/disarticulation of, 55–56, 59, 113, 301; rate of loss of, 75, 79, 80t, 81t, 84, 86; removal of, 48, 56, 193, 271–273, 288. See also cut marks on bone fertility, 143, 146, 206, 209; metaphors of, 61 fetuses, 273 (273f ), 288 fingers, 46, 47f, 50 (50f ), 129, 190, 199, 288. See also cut marks on bone

fluvial activity, 67, 85, 87. See also ENSO (El Niño Southern Oscillation) events fly larvae: intentional infestation of, 206; puparia, 37–38, 70, 72, 129, 199 foreigners/outsiders, 178–180, 204, 207–210, 216, 259, 374 forensics, 32, 73, 84, 99, 372–373 Fortes, Meyer, 172–174 foundation/construction sacrifices, 350, 355–357; of children, 14; defined, 3; at Huaca Las Ventanas, 129; at Huaca Santa Clara, 350–354 (351f, 352f, 354f ); at Kotosh, 356; at La Galgada, 356; at Templo de la Piedra Sagrada, 226; of women at Huaca Colorada, 19, 141. See also dedicatory offerings “funerary attendants,” 376. See also retainer/companion killing Galindo, 10, 147, 175 Gallinazo culture. See Virú/Gallinazo culture gender roles in sacrifice, 18f, 31, 92, 246, 362, 367, 369 genital removal/mutilation, 30, 48, 55–56, 61 geographic isotope ratios, 327 gift economy, 323, 325, 339, 366 Girard, René, 2 GMT (Grand Minimum Total), 74 gold, 123, 125, 276, 278f grave goods, 91, 96, 299, 303–305; bones as, 91, 96; crisoles, 299 Gretzer (C. T. Wilhelm) Textile Collections, 125–126 Guadalupito, 107, 108 guanaco. See camelids “guardians” of tombs, 271, 274, 288, 297, 298 hair: cutting of victim’s hair at “heat” of, 378; and Inka sacrificial practices, 13, 208–209; isotopic data

4 5 8   Index at Huaca de los Sacrificios, 185, 195–196, 208–209 (209f ); isotopic data at Huancaco (llamas), 330, 333–334 (334f ), 336–338; offerings at Llullaillaco, 209 handedness, 20, 21f, 34, 57–58 (57t), 61 hands: binding of, 37, 102, 111, 113 (113f ), 190, 199, 252–253 (253f ), 263; on disembodied forearms, 31; mutilation of, 55–56, 59; objects in, 111, 114f, 115, 125–126 (126f ), 190; positioning of, 111, 137, 138f, 165, 190, 199, 306; rate of loss of, 75, 79, 80t, 81t, 84, 86; removal of, 46, 193. See also cut marks on bone headless burials, 19, 125, 129, 131f, 356 healed injuries/lesions, 374; CA-3, 163; Chotuna-Chornancap, 184f, 186; El Brujo, 245f; Huambacho, 110, 111; Pacatnamú, 259; Plazas 3A, 3C, 37–38, 254–255; Punta Lobos, 260; TPS, 229–231, 233–234 (233f, 234f ), 237–238 healers punished for failure, 103 heart, removal of, 100, 219 herder/hunter distinction, 171, 365–366 household daily rituals, 347 huaca, defined, 24n1 Huaca Botija, 6f, 148 Huaca Cao Viejo, 6f, 19–20, 30, 91, 170t, 270t, 275 Huaca Chornancap. See ChotunaChornancap archeological complex Huaca Chotuna. See ChotunaChornancap archeological complex Huaca Colorada, 19, 141 Huaca de la Cruz, 6f, 97, 276, 287 Huaca de la Luna, 6f, 9, 17 (17f ), 34–35 (35f ), 97; broken prisoner vessels, 253 (253f ); cut marks, 40t–43t, 52f, 88, 252 (252f ); description of, 67–70 (68f ); missing bones from, 84–85; peri- and postmortem

treatment, 59–60 (59t); Platform III, 19, 68; sacrificial victims at, 249; similarities to Pacatnamú, 263–264; throat-slitting at, 252 (252f ). See also Huacas de Moche (Pyramids of Moche) Huaca de la Luna, Plaza 3A, 19, 36f, 67–70 (68f, 69f ), 76–83 (76t–77t); age, sex of victims, 38, 68; anatomical mapping of disarticulation patterns, 82t; CE, 77–78 (78t); compared to Pacatnamú, 67, 76t–78t, 80t–83t, 95, 263–264; compared to Plaza 3C, 58–60 (58f ), 62, 255; compared to Punta Lobos, 261, 263– 264; cut marks, 51–54 (50f, 52f, 54t), 68, 252, 263; data collection and analytical methods, 38–39; dedicatory burial at, 140–141; DNI (derived number of individuals), 74, 77–79 (78t), 80t; and El Niño rains, 254; human remains excavated from, 37–38; initial excavations of, 17–19, 35, 250f; LI, 77–78 (78t); MNI, 37, 76–79 (76t–78t), 80t; peri-mortem injuries, 50f, 53f, 68, 263; prisoner vessels, 253; proof from skeletonized remains, 292; rates of bone loss, 81t; sacrifice victim positions, 69f, 70, 250f; setting of, 68 (68f ); SS2 and DS2 layers, 92, 93f; taphonomic questions regarding, 65, 70, 84–86. See also Huacas de Moche (Pyramids of Moche) Huaca de la Luna, Plaza 3C, 19, 36f, 37, 44f, 45f; age, sex of victims, 37; antemortem injuries indicating warriors, 254–255; compared to Plaza 3A, 58–60 (58f ), 62, 255; compared to Punta Lobos, 261, 263–264; cut marks, 39–48 (40t– 43t, 44f, 45f, 47f ), 51–54 (52f, 53f, 54t), 102, 252 (252f ); initial excavations of, 37; modified skulls, 90–91, 92f; perimortem injuries, 37, 61; prisoner vessels, 253 (253f ); rates

Index  4 5 9 of bone loss, 81t; sacrifice victim positions, 251f; victim treatment in, 54–58 (55f ). See also Huacas de Moche (Pyramids of Moche) Huaca de los Sacrificios, 6f, 190–195 (191f–194f ); age and sex of victims, 195, 202f; data and analysis, 195–197 (196f, 197f ). See also Chotuna-Chornancap Huaca del Sol, 6f, 9 Huaca Dos Cabezas, 19 Huaca El Corte, 6f, 127, 143 Huaca El Pueblo-Úcupe, 6f, 273, 288 Huaca Fortaleza, 141 Huaca La Merced, 6f, 127 Huaca Larga, 180, 285 Huaca Las Balsas, 217 Huaca Las Ventanas, 6f, 125, 126f, 127–129 (128f ), 136 (136f ) Huaca Lercanlech, 6f; probable sacrifices in temple-top floors, 127, 128f, 141 Huaca Loro, 6f, 276–279 (277f, 278f ), 282f, 285–286; commoner burial, 123f, 124; dedicatory burial, 127; depiction of decapitation, 125–126; ENSO-related burials, 129; mountain–huaca–water triad, 143, 144f; probable sacrifices in temple-top floors, 127, 128f, 141; retainer burial, 127–129, 130f Huaca Negra (Virú valley), 326 Huaca Norte. See Huaca de los Sacrificios Huaca Rajada, 153, 271–273 (271f, 272f ) Huaca Santa Clara (Virú valley), 6f, 343–345, 353; camelid sacrifices, 326; canchero burials at, 351 (352f, 354f ); propitiatory ritual at, 350–354 (351f, 352f, 354f ); Room A-102, 346– 350 (346f, 347f, 348f, 349f, 350f ) Huacas Cao Viejo, 6f Huacas de Moche (Pyramids of Moche), 34, 70–72, 91, 170t, 249– 255 (250f–253f ), 270t, 275–276, 356.

See also Huaca de la Luna; Huaca del Sol Huaca Sialupe, 6f, 129–131, 132f, 145 Huaca Sontillo, 127, 180 Huambacho, 6f, 109–115 (110f, 112f, 113f, 114f ) Huancaco (Virú valley), 6f; analysis of camelid remains at, 329–335 (329f, 331f, 332t, 333f, 334f ); camelid sacrifice at, 319–320, 335–337 Huanchaco/Huanchaquito (Moche valley), 6f, 20, 153, 170t, 326 Huarmey valley, 238, 259. See also Punta Lobos Huayna Capac, 267, 289 Hubert, Henri, 2, 292, 364, 366–367, 378 human sacrifice: defining, 292; shift from animal to, 169–176; as subset of sacrifice rituals, 342–343; themes of at Sicán, 139–140 hydrogen isotope analysis, 340 iconography: Chimú, 12; Cupisnique, 15, 16f; decline in, 139; Nasca, 356; Sicán, 143. See also Moche culture identity of victims, 212; aclla, 268, 279; bioarchaeology and, 179, 269, 280; captive warriors, 245, 257–261, 367; at Huaca de los Sacrificios, 204; retainers, 285–287; at Túcume, 215, 220 ideology, 10, 95, 106, 118, 151, 169, 172, 176 Illimo, 6f, 277–278 infectious disease, 183, 186, 224 Initial Period warfare, 247, 248f Inka: defeat of Chimú, 12, 180, 261; retainer/companion killing, 208, 279; scholarly research on, 13–14; selection of sacrifice victims, 208. See also Chotuna-Chornancap; Huaca de los Sacrificios; Templo de la Piedra Sagrada (TPS), Túcume insects, and human corpses, 167, 199–200, 257. See also fly larvae

4 6 0  Index “invisible” sacrifices, 140 isotope ratios in plants, 327–328, 335 isotopic analysis of camelids, 321, 330–337 (332t, 333f, 334f ), 339–340 isotopic analysis of humans, 242, 246, 327; Colonial Muchik, 204; Huaca de los Sacrificios, 185, 195–196, 204, 208; Inka sacrifices, 13, 208; Pacatnamú, 259; Santa Rita B, 167 Jequetepeque valley: Chimú conquest of, 72, 259; Cupisnique society, 8, 15, 16f, 24, 104, 245f, 247, 248f; Dos Cabezas, 6f, 97, 153, 274–275, 292; female sacrifices in, 206–207; Huaca Colorada, 19, 141; social status of burials, 236–237. See also Pacatnamú; San José de Moro joint ratio scores, 75, 80; Huaca de la Luna Plaza 3A, 82t; Pacatnamú, 83t Jotoro, 180 kamayu, 207 kaolin clay ceramics, 352f, 353, 354 kero/flaring cups, 125 kinship analysis, 280, 281t, 369, 286. See also biodistance analysis Kotosh, 23 La Caleta de San José, 179–180 La Leche valley. See Túcume Lambayeque Valley Complex, 6f, 121, 178, 271–273 (271f, 272f ); Túcume, 216, 218f La Pava, 6f, 180 Larco Hoyle, Rafael, 353, 358n1 Las Avispas, 279, 285, 287 Las Casas, Bartolome de, 240 Late Horizon: capacocha/capac hucha sacrifices, 174, 240; captives/prisoners, 261–262; human and animal sacrifices, 153; identity of human sacrifices, 169; Inka camelid herding, 320; Inka human sacrifices, 239–241; mortuary patterns, 170t; Templo de la Piedra Sagrada, 220

Late Intermediate Period, 145, 170t, 178, 236, 244; camelid sacrifice during, 326; Cerro Cerrillos seated burial, 357; Chimú-style vessels, 109; overview, 255–257; representational art in, 20. See also Farfán; Huancaco; Huanchaco/Huanchaquito (Moche Valley); Pacatnamú; Punta Lobos; Templo de la Piedra Sagrada (TPS), Túcume Late Middle Sicán, 134–138 (135f, 136f, 138f ) Late Moche–Transitional Period, 174–176 Late Sicán, 11, 20, 72, 182, 278 La Viña workshop, 180 LI (Lincoln Index), 73 ligatures/bindings, 37, 115, 118–119; in art, 104–106 (104f–106f ), 115; associated with fox-serpent, 103–104; bound victims eaten by birds, 103; in Huambacho graves, 109, 111, 113–115 (113f, 114f ), 117; inferring live burials from, 139–140; on jars, 102– 103; in Moche art, 102, 252, 256; Pachacamac women, 101; representing captives, 252, 263; and retainer sacrifice, 284; at San José de Moro, 284; strangulation by, 98–101, 287; symbolic meaning of, 115, 117 “like-with-like” pattern, 152–153, 169, 170t, 200, 374 Lincoln Index (LI), 73 lips, 45f, 55–56, 61, 278 llamas, 320–322, 341n2; in columnar boxes, 141; “decapitated” ceramic heads, 136; fetal sacrifices, 131; hearts confused with children’s, 208; at Templo de la Sagrada (Túcume), 217, 220, 240. See also camelids Llullaillaco, 209 lo andino presumptions, 14, 22, 142 long bones: cut marks on, 33, 59; disarticulated/fractured, 70, 232, 257, 298; loss/removal/reburial of,

Index  4 6 1 79, 89, 95, 153; manipulation of, 89; positioning of, 190; and sample count, age, sex, 76, 79, 155, 157, 183, 223; and stature, 225; trophy-taking, 95, 262 looted items: funerary ceramics, 149n1; grave goods, 270; painted textiles, 125–126; silver kero, 125; stirrup-spout vessel, 20 looted sites: burials and cemeteries, 247, 257; Chan Chan, 279; Huaca Chotuna, 209; Huaca Las Ventanas South Tomb, 125, 126f; Morro de Eten, 15; Túcume, 216 looting, 20, 125–126, 161, 209, 216, 247, 257, 270, 279 Lord of Úcupe, tomb of, 273 (273f ), 288 maize beer (chicha), 135, 267, 295, 296, 321 “male warrior” social role, 367, 374 mallqui, 143, 206 material culture traditions, 97, 108, 146, 216, 329 Matrix 101 (Sicán) mass human sacrifice, 134–138 (135f, 136f, 138f ), 142 Mauss, Marcel, 2, 292, 364, 366–367, 378 Maya, 91, 222, 246, 375, 377–378 McClelland, Donna, 17f–18f, 30, 103, 104f Mesoamerican perspectives, 371–379 metal, 143, 207, 249, 270, 278; bracelet, 306; bronze, 11, 123–124, 127, 217; gold, 123, 125, 276, 278f; metal knife cut marks, 53f, 56, 62, 133, 188, 198; metal versus lithic tools, 31, 33, 56; sheet-metal, 298, 300; silver, 123, 125, 205, 243n3, 306; tumbaga alloy, 123 metamorphosis of dead, 378 Middle Horizon, 145, 170t. See also Late Intermediate Period; Punta Lobos Middle Sicán culture/period, 10–11,

149n1; biodistance analysis of tombs, 280–284 (281t); biodistance cluster analysis, 283f; complexities of identity, 285–287; dedicatory sacrifices, 127–129, 356–357; elite burials, 123; evidence of ritual killing during, 126–138 (128f, 130f–133f, 135f–136f, 138f ); Huaca de la Cruz during, 276–279; human sacrifice at Sicán Precinct, 139–140; human sacrifice outside of Sicán Precinct, 143–144; “invisible” sacrifices, 140; overview, 120–126; policies toward locals, 147; sacrifice in diachronic perspective, 140–142; sacrificerelated art, 124–126 (126f ); Sicán deity, 124–125, 126f, 134, 135–136 (136f ), 146; symbolic associations at Sicán Precinct, 142–143; torching of temples, 11; typical commoner burial, 123f, 124. See also Huaca Loro; Lambayeque Valley Complex; Late Sicán military roles and ritual life, 369 missing body parts, 373; bones missing in burials, 84–87; ceramics of, 30–31; dismemberment, 46–48, 50; feet missing in burials, 271–273, 288; reasons for, 37, 79, 80t, 81t, 84 mitochondrial DNA (mtDNA), 123– 124, 254, 285 MLNI (most likely number of individuals), 73–74; Huaca de la Luna Plaza 3A, 77–78 (78t); Huaca Las Avispas, 279; Pacatnamú, 76–79 (76t–79t), 81t; Templo de la Piedra Sagrada, 226 MNI (minimum number of individuals), 73–74; Chimú burial, 279; Huaca de la Luna Plaza 3A, 37, 76–79 (77t–78t), 80t; Huaca de la Luna Plaza 3C, 37; Pacatnamú, 76–79 (77t–79t), 81t Moche culture, 15–17, 324; accuracy of sacrifice art, 17, 29, 60–62, 101–102; captives in art, 101, 244; chronology

4 6 2   Index of, 9–10; hunting of deer, sea lions in art, 324; no mutilation of children, 145; ocean and mountains in art, 143, 205; ritual strangulation in art, 102–106 (104f, 105f, 106f ), 115; sacrifice in art, 101–102; source of captives, 254; Southern Moche society, 106–108; “Warrior Narrative” in, 18f, 30–31, 101, 249, 264; warriors in art, 255, 264 “Moche state religion,” 151–152 Molina, Cristóbal, 100, 207 monumental architecture, 29, 151, 154, 247, 263 morbidity and life history, determining, 183–185 (184f ) Morro de Eten, 6f, 15 Mórrope, 205; child sacrifice in 18th-century myth, 146, 205; Colonial-era isotopic variation of, 170t, 204 mortuary patterns: Andean through time, 170f; versus sacrificial practice, 95–96 mountain–huaca–water triad, 143, 144f mourners, 348–349 Muchik peoples, 10–11, 129–134, 143–147, 183 mud, 60, 88, 92 mummification, 66 musical instruments, 313, 378 Nasca, 356 Naymlap, legend of, 182 neck: captives led around by, 246; in decapitation, 49, 133, 217; hyperextension of, 58; kyphosis of, 111; ligatures around, 99, 100, 102, 104f, 105 (105f ), 106f, 108, 111, 246, 251f, 252, 253 (253f ), 263, 284, 292; necklaces, 111, 300; sash around, 276; strangulation, 98–99, 101, 111, 276, 292; of surrogate sacrifices, 102; tied to post, 113–115 (114f ). See also throat-slitting

necropampa immolation, 13, 267. See also retainer/companion killing Nectandra, 133, 241–242 Nepeña Valley, 6f, 109–115 (110f, 112f, 113f, 114f ) Nias Island (Indonesia), 356 nonliving object sacrifice, 342 North Coast of Peru, 6f, 5–12; chronology, 7–12 (7f ); natural setting, 5, 6f; ritual violence, study of, 371–375 noses: broken, 370; covering of in burial, 137, 138f; mutilation/removal of, 45t, 56, 61; ornaments, 246, 275 object sacrifice/immolation, 14–15, 103, 136 (136f ), 342, 358, 363 oral/dental health, 134, 138, 184 (184f ), 186–187, 195, 203f, 204, 286, 230–231 osteoarthritis, 124, 138 oxygen isotope analysis, 185, 242, 340 Pacatnamú, 6f, 20, 258f; age and sex of victims, 257, 260, 263; antemortem injuries indicating warriors, 263–264; captive warrior evidence, 257–259; CE, 77–78 (78t); compared to Huaca de la Luna, 65, 76t–78t, 80t–83t, 95, 263–264; cut mark data from, 87–88; dedicatory sacrifices, 357; DNI (derived number of individuals), 74, 77–79 (78t), 81t; evidence of chest-opening, 72; intentional exposure of remains, 262–263; LI, 77–78 (78t); missing bones from, 86–87; perimortem trauma to crania, 85f; preserved rope around ankles, 258f; sacrifice victim positions, 71f; similarities to Huaca de la Luna, 263–264; taphonomic winnowing of, 86–88 (87f ) Pachacamac, 12, 100–101, 124–125, 287 paleopathological data, 221–223, 236, 372

Index  4 6 3 paleteada pottery, 180 Paloma, 14 Pampa de Burros, 180 Pampa de Chaparrí, 180 Pampa Grande, 6f, 121, 141 Pañamarca, 6f, 31, 105, 105f, 109 parry fractures, 164–165 (164f ), 195, 255, 261 pateras-like vessels, 353 penis, 30, 48, 55–56, 61 perimortem trauma, 86, 163–165 (164f ), 168, 185, 195, 197f, 198, 217, 231, 284, 368; CA-3, 163; El Castillo, 19, 108, 117; Huaca de la Luna, 38–39, 50, 68, 263; Pacatnamú, 72, 85f, 86; Punta Lobos, 260; San José de Moro, 308–309; Santa Rita B (Chao valley), 158–160, 165, 168 periosteal infections/lesions, 186, 195 personhood, 22, 27, 94, 139–140, 288 Piedras Negras, 356 pit tombs, 295, 296 Pizzaro, Pedro, 267 “place-making,” 355 plants: isotope ratios in, 327–328, 335; metaphors of, 15, 125, 206 poisoning, 139, 287, 373 polisemy, 378 politics: of animal sacrifice, 324, 339; and captive killings, 245, 256, 264; Chimú, 179; cohesion through sacrifice, 173–176; of conquest, 207–208; of ethnicity, 147; foundation sacrifices, 357; integration and control through violence, 4; Late Sicán, 278; Middle Moche, 10; Middle Sicán, 121–122, 147; Moche, 9–10, 106–107, 151–152, 264; of Muchik blood sacrifice, 145–147; Nepeña, 118; of prestige sacrifice, 173, 267–268, 366; and ritual “third space,” 22; and structuring effects of sacrifice, 362, 368–369. See also ENSO events porotic hyperostosis: at ChotunaChornancap, 183, 184f, 186, 195,

224; at Túcume, 228–230 (228f, 229f, 230t), 237 postmortem treatment: Huaca de la Luna, 59–60 (59t); Mesoamerica, 378 “potentially sacrificed individual,” 140, 287 Practice Theory, 362 pregnancy, 146, 206, 273, 276, 277f, 284, 288 premortem, 56 “preprogrammed repetitive action” of sacrifice, 100, 364–365 Presentation Theme. See Sacrifice Ceremony/Presentation Theme prestige sacrifices, 173, 267, 366 priestess burials: Priestess of Chornancap, 278; San José de Moro (various), 274; San José de Moro M-U1221, 303–311 (304t, 304f, 305f, 307t, 308f, 310f, 311f ); San José de Moro M-U1525, 299–303 (300t, 301f ) priestess in Sacrifice Ceremony, 18f, 31, 92, 298 prisoners. See captives/prisoners “problematic deposits,” 376 prolonged burial/delayed interment, 129, 201, 288. See also re-interment of body parts/secondary interments; reopening of graves propitiatory ritual, 317, 342–343, 356; as calculated sacrifice, 172–173; Huaca Santa Clara, 350–354 (351f, 352f, 354f ); substitute sacrifices in, 240. See also dedicatory offerings; foundation/construction sacrifices Punta Lobos, 6f, 230t; age and sex of victims, 20, 237, 260, 263; cranial deformation patterns at, 235t, 238; exposure of bodies, 263; on isolated hillside, 260–261; killing of captives, 257, 259–261 (258f, 260f ); lack of antemortem trauma, 263–264

4 6 4   Index Pyramids of Moche, 34, 70–72, 91, 170t, 249–255 (250f–253f ), 270t, 275–276, 356. See also Huaca de la Luna; Huaca del Sol Qhapac hucha (capacocha/capac hucha) sacrifices, 13, 100, 153, 174, 209, 240–241 Quilter, Jeffrey, 10, 171, 356 radiocarbon/metric dating, 20, 37, 152, 155t, 182, 216, 257, 261, 320, 343 rainfall. See ENSO (El Niño Southern Oscillation) events rates of loss for skeletal elements, 79; Huaca de la Luna Plaza 3A, 80t; Pacatnamú, 81t reciprocal obligation theories, 142 Reinhard, Johan, 100, 207–208 re-interment of body parts/secondary interments, 152–153, 165, 169, 170t, 193, 374 religion: and captive killings, 262; Chimú, 20; Cupisnique, 8, 15; Early Moche, 9; and El Niño events, 138; Inka, 325, 357; Late Sicán, 11; Middle Moche/Huaca de la Luna, 9–10, 35, 222; Middle Sicán, 10, 120–122 (122f ), 125, 127, 134, 147; “Moche state religion,” 151; Muchik ritual, 147; overlapping with politics, 339; permeation of rituals in, 153–154; prestige sacrifices, 173; and ritual strangulation, 118; and role of sacrifice, 362, 363–364, 366–369, 374– 377; sacrificial substitution, 172; in Santa/Nepeña regions, 116–118; and secular punishment, 22; shared Moche, 106–108; of subject populations, 207–208, 210, 220; women’s role in, 207 reopening of graves, 292, 298, 306; at Huaca Cao Viejo, 275; at Huaca de los Sacrificios, 193; at Huambacho, 117; at Matrix 101, 135, 137; at M-U1221 burial, 306, 311, 314;

at M-U1525 burial, 299, 301–303, 311; re-interment of body parts/ secondary interments, 152–153, 165, 169, 170t, 193, 374; at San José de Moro, 274; at Santa Rita B, 171. See also prolonged burial/delayed interment reprisal/retaliatory killings, 261, 262–264 retainer/companion killing, 3, 170t, 269t–270t, 349–350 (349f ), 369, 374, 376; acompanante, 266; age and sex of victims, 128, 139, 285– 289; bioarchaeology of, 268; biodistance analysis, 280–284 (281t, 282f, 283f ); Chimú and Inka Periods, 279; compared to TPS, 239–240; evidence for sacrifice, 287–289; evidence of, 269, 269t–270t; identities of retainers, 125, 142–143, 239, 267–268, 285–287, 369; Middle Sicán Era, 127–129, 276–279 (277f, 278f ); Moche Era, 270–276 (271f, 272f, 273f ), 297; noncontemporaneous deaths, 3, 143, 288–289; physical evidence of violence, 284–285; San José de Moro, 297, 298; in West Mexico, 377. See also “funerary attendants” retaliatory killings, 261, 262–264 revenge killing, 367–368 ritual: bleeding/killing of animals, 101, 322–324; canchero burials, 351 (352f, 354f ); Chotuna-Chornancap killings, 200–203 (201f–203f ); elite combat, 254; household daily rituals, 347; human sacrifice, 342–343; killings in Middle Sicán culture/period, 126–138 (128f, 130f–133f, 135f–136f, 138f ); killings of animals, 365–366; military roles and, 369; and nonritual violence, 369–370; power through, 4, 176, 238; scholarship on human ritualized violence, 1–5, 371–379; strangulation, 102–109 (104f, 105f,

Index  4 6 5 106f ), 115; tribute versus captive executions, 367. See also propitiatory ritual ritual killings, classification of, 3–4 Ritual Sacrifice in Ancient Peru (Benson & Cook), 5 ropes. See ligatures/bindings Rowe, John H., 182, 262 Rubiños y Andrade, Justo Modesto, 146, 182, 205 runa sacrifice, 13, 240–241 “sacred heart matter,” 378 sacrifice: of animals, 174, 322–324; calculated, 172–173; complex nature of, 4; definition of, 2–3, 100, 292, 363–366; demonstrating local authority, 173; drugging of victims, 241–242; emic significance of, 3–4; versus killing, 322–324; modeling Andean, 239–242; as “overkill,” 365; prestige, 173; as a process, 2; as religious practice, 362; rites to transform, 366–367; role of, 374; and ruling elite, 4–5; at Sicán capital, 127–129 (128f ); sociopolitical cohesion through, 173–174 “sacrifice as offering,” 142 Sacrifice Ceremony/Presentation Theme, 17, 30–31, 219, 298; binding in, 102, 104–105 (104f, 105f ); disarticulation in, 91–92; possible heart extraction in, 61–62; strangulation in, 102, 105f, 118; throat-slitting in, 18f, 31, 61, 101 San José de Moro, 6f, 97, 274, 284; as ceremonial and pilgrimage center, 296; cultural chronology of, 294f; funerary practices at, 296–298; M-U615 burial, 174; M-U1221 burial, 303–311 (304t, 304f, 305f, 307t, 308f, 310f, 311f ); M-U1525 burial, 299–303 (300t, 301f ); M-U1525 priestess burials, 299–303 (300t, 301f ); perimortem injuries, 308–309; periods of

occupation, 293–296; possibility of human sacrifice, 311–313; postmortem treatment, 309–310. See also Jequetepeque valley San José Mogote, 356 Santa Rita B (Chao valley), 6f; age and sex of victims, 155–158; analysis of human bone from, 155–163 (155f, 162f, 163f ); archaeological context, 154–155; burial positions, 165–167 (166f ); camelid sacrifice, 152, 154–155, 158f, 161, 167–168 (168f ), 171–176, 326; perimortem injuries, 158–160, 163–165 (164f ) Santa Valley: ritual strangulation in, 107, 108. See also El Castillo de Santa scanning electron microscope (SEM), 34, 38–39, 52f scholarship on human ritualized violence, 2–5, 371, 375 sea lions, 101, 324 secondary interments, 153. See also reopening of graves seed metaphors, 206 Señora de Cao, 275 sex of victims. See age and sex of victims sharp force trauma, 217–219 (219f ); at Cerro Cerrillos, 133f; in chest opening, 197f, 198 (198f ); at Chornancap, 188, 189f; in decapitation, 32, 56, 188, 189f; discerning behavior and intent from, 32–34; in dismemberment, 50f, 56, 86; at Huaca de la Luna Plaza 3A, 49–50 (50f ), 86; at Huaca de la Luna Plaza 3C, 39, 56; at Huaca de los Sacrificios, 197f, 198–199; in throat-slitting, 39, 49, 50f, 56, 133 (133f ); at TPS, 239–240. See also chest-opening; decapitation; throat-slitting shrouds, 107, 110, 113, 145, 199, 217, 265, 288, 348 Sicán (Lambayeque) culture. See Middle Sicán culture/period

4 6 6   Index Sicán religious precinct, 121–122 (122f ), 148 silver, 123, 125, 205, 243n3, 306 Sipán, 6f, 97, 271–273 (271f, 272f ) skeletal trauma, 158–160, 164f, 165 skeleton, role of, 61 “skull-seed,” 378 social hierarchy. See elites social identity, 236–238, 242–243, 369 sociopolitical integration, 4 Southern North Coast of Peru, 7f Spondylus shells, 13, 200; at Cerro Cerrillos, 133; at Chornancap Norte, 187f, 190; at Chotuna-Chornancap, 205; force feeding at Huancaco, 330 (331f ); at Huaca de los Sacrificios, 200–201; at Huambacho, 111; at Punta Lobos, 259; in sacrificed llamas, 330, 331f; at Sicán, 127; at Túcume, 205, 217 stone reliefs, 15, 248f strangulation, 100–102, 292; after excarnation, 117; forensic evidence for, 99; Huaca de la Cruz, 276; in Moche art, 102–106 (104f, 105f, 106f ); “nonspectacle” aspect of, 106, 115–116, 365, 369; physical anthropology of, 98–100, 287; Sacrifice Ceremony, 102, 105f, 118; snakes and, 103–106 (104f, 105f, 106f ); symbolic meaning of ligatures, 115, 117 stratigraphic analysis: Chornancap Norte, 202–203; effect of earthworms on, 65; Huaca Cao Viejo, 275; Huaca de los Sacrificios, 192–193, 203; Huaca Loro West Cemetery, 129; Matrix 101, 72–73; Pacatnamú mass burial, 72; Plaza 3A, 68–70, 91–92, 93f; Plaza 3C, 37, 60; Santa Rita B, 168 strontium isotope analysis, 167, 242, 340 structuring effects of sacrifice, 362 suicide/auto-sacrifice, 3, 15, 289

surrogates, 102, 354 symbolic inversion, 98, 101, 109–110, 115–118, 129, 365 symbolic violation, 116–117, 312, 314 Tambo Real workshop, 180 taphonomy: defined, 65–66; in mass burial sites, 65–67; methods, 72–75, 372; questions regarding Huaca de la Luna, Plaza 3A, 70; questions regarding Pacatnamú, 65, 72; regarding non-adult bone, 159–160; of sacrifice victims, 88–95 (89f, 90f, 92f, 93f ) tattoos, 348 Tel Haror (Israel), 175 Tello, Julio C., 125 Tell Umm el-Marra (Syria), 176 Templo de la Piedra Sagrada (TPS), Túcume, 6f, 218f; age and sex of victims, 223, 226f, 227–228 (227f ), 237, 241; camelid remains, 217, 325; dedicatory offerings at, 223; description/chronology of site, 216–220 (218f, 219f ); foundation/ construction sacrifices at, 226; lack of grave goods, 217; Late Horizon, 220; MLNI (most likely number of individuals), 226; sharp force trauma at, 239–240; weaving women, 279. See also Túcume thoracic: cavity/cage, 133, 188, 190, 198f, 199–200; vertebrae, cut marks on, 32, 40t, 42t, 44, 46, 188, 259, 260f throat-slitting, 20, 265; among Muchik people, 144–145; autosacrifice, 15; of captives, 246, 252; Cerro Cerrillos, 133 (133f ); Chimú vessel depicting, 20, 21f; Chornancap Norte, 188, 189f, 201–202 (201f ); cut mark patterns in, 32; directionality in, 53, 54t, 57–58 (57t); Huaca de la Luna Plaza 3A, 49, 50f, 53 (54f ), 59f, 68, 101; Huaca de la Luna Plaza 3C, 39,

Index  4 67 44f, 53 (54f ), 55–56, 59f; Huaca de la Luna/Pyramids of Moche, 252 (252f ); Huaca de los Sacrificios, 198–199, 201 (201f ); Huaca Las Ventanas, 129; Matrix 101, 137; in Moche art, 18f, 61–62; Pacatnamú, 257; Punta Lobos, 259–260 (260f ); reported of Inkas, 207–208; respectful burial after, 265; in Sacrifice Ceremony, 18f, 31, 61, 101; symbolism of, 115; TPS (Túcume), 217–219 (219f ), 239–240; use of tumi for, 56. See also blood sacrifice; decapitation Tiwanaku, 356 “tomb choreography,” 268 tombs. See entombment tools: metal versus lithic, 31, 33–34, 56; sacrificial, 105; types of, 56; weaving, 279. See also handedness trauma, 85f; blunt force, 88; skeletal trauma, 19, 164f, 165. See also cranial trauma; perimortem trauma; sharp force trauma trophy collection, 94, 125–126 (126f ), 368–369, 378 Túcume, 20; as Chimú and Inka capitals, 179, 180, 216; compared to Andean models, 239–242; health, identity, and social status, 236–238; Huaca Larga, Huaca I, Huaca Las Balsas, 217; reestablished capital at, 134; sacrificial versus cemetery remains at, 223, 226–236 (226t, 227f–229f, 230t, 232t, 233f–234f, 236f ); site and chronology, 216. See also Templo de la Piedra Sagrada (TPS) tumbaga alloy, 123 tumi (crescent-bladed) knives, 30–31, 56, 103, 105, 125–126 (126f ), 278 Turkey, 175 turquoise pieces in sacrificed llamas, 330, 331f Tylor, E. B., 2, 323

Úcupe, 141, 180, 273 (273f), 288 Uhle, Max, 12, 19, 100–101 ulluchu fruit, 105, 106f urban zones, 322; Huacas de Moche, 9, 35, 68, 91, 276; Pampa Grande, 10, 121, 141; Santa Rita B, 154; Túcume, 216; USUIs (Urban Settlements of Uncertain Degrees of Incorporation), 151, 153–154, 169–170 Vanuatu (Melanesia), 321 Ventarrón, 6f, 15, 23 vicuña. See camelids violence: ritual and nonritual, 369–370; and the sacred, 2; in US society, 370 Virú/Gallinazo culture, 8, 358n1; Gallinazo Group, 6f, 329, 343, 347, 354, 358n1; Huaca Santa Clara, 325–326, 343; relationship to Moche, 8, 9, 19, 107–109, 329; retainer/companion killing, 276; ritual strangulation within, 108. See also camelids warfare/warriors, 264; antemortem injuries indicating, 239, 263–264; bioarchaeological evidence of, 37, 225, 239–241, 247–250, 255, 257–258; Cerro Sechín depictions, 15, 248f; on Huaca Loro textile, 126; “male warrior” social role, 367, 374; Moche “Warrior Narrative,” 18f, 30–31, 249, 264; Punta Lobos not involving, 260–261, 263–264; runa killings, 13; stone reliefs depicting, 248f; Warrior-Priests, 272, 276. See also captive executions; Sacrifice Ceremony/Presentation Theme Wari and Moche, 124, 295, 368–370 water and other liquids: canals/channels for, 143, 205, 207; kamayu and running water, 207; meaning of to Muchik, 145–146; mountain– huaca–water triad, 143, 144f; and transport of bones, 85, 87, 305

4 6 8   Index women: birth/pregnancy, 206, 273, 276, 277f, 284, 288; in blood sacrifice, 204; careful burials of, 264; dedicatory offerings of, 357; foundation sacrifices, 19, 141; in M-U1221/ U1525 burials, 302–303, 305–306, 311–312; nonritual violence against, 370; Pachacamac capacocha sacrifice, 100–101; participants in captive sacrifice, 246; in pit tombs, 296; Priestess of Chornancap, 278–279;

shorn hair in sacrifices, 208–209 (209f ); weaving women at Túcume, 268, 279. See also retainer/companion killing Yassıhöyük, 175 Zaña valley, 14, 273 (273f ) Zárate, Agustín de, 267–268 zooarchaeology, 316, 324 Zuidema, R. Tom, 182