Tibes : People, Power, and Ritual at the Center of the Cosmos [1 ed.] 9780817382520, 9780817316860

The first comprehensive analysis of a strategically located ceremonial center on the island of Puerto Rico. The prehisto

157 17 4MB

English Pages 349 Year 2009

Report DMCA / Copyright

DOWNLOAD PDF FILE

Recommend Papers

Tibes : People, Power, and Ritual at the Center of the Cosmos [1 ed.]
 9780817382520, 9780817316860

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

Tibes

CAR IBBE AN ARCHAEOLOGY AN D ETHNOHISTORY Series Editor L. Antonio Curet

Tibes People, Power, and Ritual at the Center of the Cosmos

Edited by L. AN TON IO C U RET AN D LISA M. ST R I NGER

THE UNIVERSITY OF ALABAMA PRESS Tuscaloosa

Copyright © 2010 The University of Alabama Press Tuscaloosa, Alabama 35487-0380 All rights reserved Manufactured in the United States of America Typeface: Minion ∞ The paper on which this book is printed meets the minimum requirements of American National Standard for Information Sciences-Permanence of Paper for Printed Library Materials, ANSI Z39.48-1984. Library of Congress Cataloging-in-Publication Data Tibes : people, power, and ritual at the center of the cosmos / edited by L. Antonio Curet and Lisa M. Stringer. p. cm. — (Caribbean archaeology and ethnohistory) Includes bibliographical references and index. ISBN 978-0-8173-1686-0 (cloth : alk. paper) — ISBN 978-0-8173-5579-1 (paper : alk. paper) — ISBN 978-0-8173-8252-0 (electronic) 1. Tibes (Ponce, P.R.)—Antiquities. 2. Indians of the West Indies—Puerto Rico—Tibes (Ponce)—Antiquities. 3. Centro Ceremonial Indígena de Tibes (Tibes, Ponce, P.R.) 4. Excavations (Archaeology)—Puerto Rico—Tibes (Ponce) 5. Indians of the West Indies—Puerto Rico—Tibes (Ponce)—Rites and ceremonies. 6. Sacred space—Puerto Rico—Tibes (Ponce)—History. 7. Tibes (Ponce, P.R.)—Social life and customs. 8. Landscape—Social aspects—Puerto Rico—Tibes (Ponce)—History. 9. Power (Social sciences)—Puerto Rico—Tibes (Ponce)—History. 10. Social classes—Puerto Rico—Tibes (Ponce)—History. I. Curet, L. Antonio, 1960– II. Stringer, Lisa M., 1966– F1969.T53 2010 972.95'7—dc22 2009026832

To the original members of the Sociedad Guaynía de Arqueología e Historia de Ponce and the Sociedad Arqueológica del Sur-Oeste de Puerto Rico, pioneers of the archaeology of Tibes. In Memory of José Lugo, late director of the Centro Ceremonial Indígena de Tibes.

Contents

List of Figures and Tables Acknowledgments

ix

xiii

1. Introduction L. Antonio Curet and Lisa M. Stringer

1

2. Tibes: History and First Archaeological Work Pedro Alvarado Zayas and L. Antonio Curet 19 3. The Archaeological Project of the Ceremonial Center of Tibes L. Antonio Curet 38 4. Geophysical Prospection at the Ceremonial Site of Tibes, 1998–2001 Daniel Welch 60 5. Paleoethnobotanical Research at Tibes Lee A. Newsom 80 6. Animal Use at the Tibes Ceremonial Center Susan D. deFrance, Carla S. Hadden, Michelle J. LeFebvre, and Geoffrey DuChemin 115 7. Lithics from the Tibes Ceremonial Site: Analysis of the Stone Artifacts from the 1996–1999 Field Seasons Jeffery B. Walker 152 8. Boulder Lithology Survey at the Tibes Ceremonial Site Scott Rice-Snow, Melissa J. Castor, Andrew K. Castor, Jeffry D. Grigsby, Richard H. Fluegeman, and L. Antonio Curet 177

viii / Contents 9. Ancient Bones Tell Stories: Osteobiography of Human Remains from Tibes Edwin F. Crespo-Torres 191 10. Bone Chemistry and Paleodiet at the Ceremonial Center of Tibes William J. Pestle 209 11. Tibes and the Social Landscape: Integration, Interaction, and the Community Joshua M. Torres 231 12. Plazas, Bateys, and Ceremonial Centers: The Social and Cultural Context of Tibes in the Ancient History of Puerto Rico L. Antonio Curet and Joshua M. Torres 261 References Cited Contributors Index

325

287 323

List of Figures and Tables

Figures 1.1. Map of Puerto Rico showing the location of the Ceremonial Center of Tibes 9 1.2. Topographic map of the region around the Ceremonial Center of Tibes 1.3. Topographic map of the Ceremonial Center of Tibes 2.1. Photographs of the Ceremonial Center of Tibes

13 22

3.1. Map of the Ceremonial Center of Tibes showing the location of the archaeological excavations conducted by the current project 40 3.2. Ceramic concentration map of Tibes 3.3. Shell concentration map of Tibes

41 44

3.4. Chronological sequence of the deposits and clusters of burials from Tibes 45 3.5. Sketch of excavation units

47

3.6. Graphic distribution of calibrated radiocarbon dates for Tibes 4.1. Geophysical survey areas during 1998

57

63

4.2. Two depictions of the 1998 North data set of electrical resistivity

65

4.3. Three depictions of the 1998 East data set of electrical resistivity

67

4.4. Two depictions of the 1998 West data set for electrical resistivity

69

4.5. Three depictions of the 2001 North data set

73

10

x / Figures and Tables 4.6. Anomaly in Structure 9

75

4.7. Two depictions of the 2001 South data set

76

4.8. Depiction of the 2001 low-pass filtered data for the whole site

78

5.1. Section images of leguminous wood Type 41 (Fabaceae, cf. Pictetia sp.) from Tibes Feature 03-2 94 6.1. Relative abundance of the molluscan MNI from Unit N276 E105, Saladoid context, and Unit OP19E, Ostionoid context 135 6.2. Preferred habitats of animals identified in the Tibes faunal assemblage

142

6.3. Invertebrate MNI frequency by preferred habitats for mollusks present in N276 E105 and OP19E 148 7.1. Examples of lithic artifacts

160

7.2. Examples of lithic artifacts

161

8.1. Locations of two boulders sampled from the structures at Tibes

186

8.2. Locations of three possible calcareous sandstone source areas, external to the Tibes site 187 9.1. Map of Puerto Rico showing the locations of archaeological sites where human remains have been recovered 192 9.2. Artificial fronto-occipital cranial deformation

199

10.1. Values of d 15N from Tibes humans and possible elements of protein food web 223 10.2. Mean relative contributions of protein sources to diets of Tibes individuals 226 11.1. Map of Puerto Rico showing the project area 11.2. Distribution of site types through time

237

242

11.3. Settlements and cost boundaries, Period II 11.4. Period II pottery stylistic distributions

242 244

11.5. Calibrated radiocarbon dates for Period II

246

11.6. Settlements and cost boundaries, Period III

248

11.7. Period III pottery stylistic distributions

251

11.8. Calibrated radiocarbon dates for Period III 11.9. Settlements and cost boundaries for Period IV 11.10. Period IV pottery stylistic distributions

253 255 257

Figures and Tables / xi

Tables 2.1. General information on the ceremonial structures at the Ceremonial Center of Tibes 24 3.1. Radiocarbon dates obtained from four different contexts at Tibes

56

5.1. Primary arboreal taxa found in the southern dry limestone region of Puerto Rico 87 5.2. Taxonomic assignments for Tibes excavated seed assemblage

91

5.3. Taxonomic assignments for Tibes excavated wood charcoal assemblage 5.4. Distribution of seeds and other non-wood plant remains

92

96

5.5. Distribution of select wood taxa by provenience expressed as counts and ubiquity 98 5.6. Non-wood ethnobotanical uses among select taxa present in Tibes archaeobotanical assemblage 105 6.1. Habitats, zones, and aquatic habitats of the Ponce region, southern Puerto Rico 117 6.2. Analyzed contexts and cultural affiliations

119

6.3. Allometric regression formula and values used in this study

120

6.4. Species identified from the 1996–2003 excavations at the Ceremonial Center of Tibes 122 6.5. Summary of vertebrate and invertebrate NISP and MNI by context 6.6. Relative abundance of taxa by class and mesh size

127

129

6.7. MNI of molluscan taxa in Units N276 E105, Saladoid context, and OP19E, Ostionoid context 134 6.8. MNI of mollusks unique to Units N276 E105, Saladoid context, or OP19E, Ostionoid context 136 6.9. Bone and shell modifications exclusive of burning 6.10. Vertebrate fauna by preferred habitat 6.11. Invertebrate fauna by preferred habitat

137

138 146

7.1. Excavation units as originally labeled in 1998, as relabeled in 2003, and as used herein 162 8.1. Mix of boulder rock types in all surveyed Tibes pavements compared to that in the Portugués River 181

xii / Figures and Tables 8.2. Significant differences in mix of boulder rock types among surveyed pavement areas at Tibes 182 8.3. Average boulder sizes for the rock types represented in the surveyed Tibes pavements 185 9.1. Sex/age distribution in Tibes

195

9.2. Stature estimation at Tibes and other archaeological sites in Puerto Rico 195 9.3. Pathological conditions observed at Tibes and other archaeological sites in Puerto Rico 196 9.4. Age-at-death distribution observed at Tibes, Punta Candelero, and Paso del Indio archaeological sites 201 10.1. Individuals sampled for stable isotope analysis

216

10.2. Quality standards and assessment of individuals interred at Tibes

220

10.3. Analyzed stable isotope ratios of individuals interred at Tibes, predicted isotopic values of diets, and AMS dates 221 10.4. Food web groupings employed for paleodietary reconstruction

222

10.5. First percentile, mean, and 99th percentile values for multisource mixture reconstruction of Tibes’s paleodiet 225 11.1. Site information within the project area

239

Acknowledgments

The study of social change in the ancient Caribbean is a relatively young interest in the archaeology of this region, and the project that is the backbone of this volume is an example of this. Nonetheless, despite the lack of experience for conducting this kind of study, it was clear from early in the project that a joint effort among many specialists, field personnel, government officials, funding institutions, and staff members of the Centro Ceremonial Indígena de Tibes was necessary in order for it to be successful. From this perspective it has to be admitted that this volume presents only the partial fruits of these efforts, particularly the results of the work of specialists. What is not shown in these pages is all the work, assistance, and support provided by a large number of people and organizations that make up the scaffolding that is used to build solid research and educational programs. These are the unsung heroes of Tibes and to them we owe our gratitude. We would like to first thank the people of Ponce for their warm hospitality and support during the many years the project has been working in one of the most important archaeological sites of the region. Especially, we want to recognize the late Mayor Rafael “Churumba” Cordero Santiago, who for many years appreciated and supported the work we were doing both at the site and in the community. His dedication to Puerto Rican culture and its archaeological heritage was always an inspiration for us. We also are indebted, first, to Ms. Maruja Candal and, second, to Ms. Vangie Rivera, consecutive Secretaries of Culture of the Autonomous Municipality of Ponce. Their tireless work and support of our project considerably facilitated our work. However, more important, they provided the project with an essential educational component. They, their staff, and the staff of the park taught us and continuously reminded us that we do not own the archaeological record; we

xiv / Acknowledgments are just its stewards. The real owners are the people of Ponce and Puerto Rico, and the least we can do as archaeologists is return to them this heritage by doing public outreach in the form of popular publications, public lectures, workshops for children, and other activities. The staff of the Centro Ceremonial Indígena de Tibes deserves our deepest gratitude and appreciation. We do not have words to express what their warm hospitality and endless patience year after year have meant to us. Their great dedication to the park and their love toward Tibes is an inspiration to all of us. Especially, we would like to thank José Lugo, the late director of Tibes, who always gave us his unconditional support, both logistically and morally. We also would like to thank Irma Zayas Alvarado, current director; Irma Santiago Rivera, assistant director; Carmen Martínez Ajá, curator; Luis Rodríguez Gracia, former director; William Rodríguez, cashier; and the administrative assistants Carmen Lillian Pérez and Sandra Vega. We are in debt to the tour guides Carmen Martínez Alvarado, Luis Sánchez, Salvador Mas, Hector Ortíz, Milagros Robledo, Walter González, Maritza Santos, Nelson Cintrón, and Carlos Deida, who provided us with the means to reach the general public and taught us their language. They were a constant reminder that our professional duty was not only with the scientific community but also with the people of Ponce and Puerto Rico. The maintenance personnel of Tibes: Ramón Almodovar, head of maintenance; Melquiades Medina, carpenter; Reinaldo (Jalisco) Flores; Edwin Morales; Luis Martínez; and Sonia Meléndez, deserve not only our thanks for all the help they have provided us but also the credit for keeping Tibes the way it is, one of the jewels of Puerto Rico. The loyalty and love that all these people have toward Tibes is probably the most valuable resource that the park has. They are the true heroes of Tibes that make all visitors have an unforgettable educational experience and at the same time maintain and conserve the site for future generations and provide support for active archaeological research. To all of them, “muchas gracias de todo corazón.” We would also like to thank all the personnel of the Hotel Meliá for their patience and friendly treatment toward us. We really appreciated their smiling faces every afternoon when we arrived from the field all dirty and smelly. Our deepest gratitude goes especially to Nicolás Albors, president; Enrique Albors, manager; and Cathy Becerra, director of reservations, for their support and for making our lives a lot easier by providing us with a hot shower and a cool room. Ray Petty, president of the Fundación del Centro Ceremonial de Tibes, has also been a strong and avid ally, friend, and supporter of the project since its inception. We appreciate all his effort, work, and assistance throughout all the years we have been working at Tibes. We are also in debt to a number of colleagues, friends, and family who always offered their help and who in one way or another contributed directly or indirectly

Acknowledgments / xv to the project. First of all, Lee Newsom, co-director of the project from 1995 to 2003, was instrumental in many aspects of the project. She not only helped in designing the methodology to follow but also educated us in many of the intricacies of the analysis of biological samples in archaeology. Pedro Alvarado Zayas, Juan González Colón, Luis Rodríguez Gracia, Edgar Maiz, and Eduardo Questell, founding members of the Sociedad Guaynía, provided us on one occasion or another with valuable information and suggestions. Listening to their anecdotal stories of the harshness of the early years of archaeological research at Tibes did nothing but humble us and dwarf any of our efforts at the site. Many people provided us with suggestions, materials, and other support. Discussions with Miguel Rodríguez, José Oliver, Reniel Rodríguez, Joshua Torres, Edwin Crespo, and Jaime Pérez on many of the issues we are dealing with helped us to better focus and reevaluate many of the premises and goals of the project. Ramón Almodovar and Carmen Martínez Almodovar provided us with specimens of plants and animals and with their extensive knowledge of them. Their contributions to the project were essential to the studies of faunal and botanical remains and bone chemistry, the results of which are included in this volume. Melquiades Medina provided his skill as carpenter, building a variety of equipment for our fieldwork. Jill Seagard deserves credit for the figures in Chapters 1, 2, 3, 8, and 9. We are also in debt to a large number of support staff and volunteer students from the following institutions: Gettysburg College, Southern Illinois University at Carbondale, University of Colorado at Denver, University of Pennsylvania, University of Illinois at Chicago, University of Florida, and the Field Museum. Most of the studies included in this volume were made possible by the generous financial support of many institutions, including Gettysburg College, National Geographic Society (Grants #6260-98 and #7276-02), Heinz Foundation, National Science Foundation (DUE-#9551495, Ref. #0106520), the Grainger Research Fund of the Field Museum, and donations from Ms. Beverly Gunness, Mr. and Mrs. William S. Macdonald, Mr. and Mrs. William A. Osborn, Mr. and Mrs. Richard J. Pigott, and Mr. and Mrs. Edward J. Zander. We want to thank our friends and family who stood by us throughout the length of the project. Jaime Pérez, a friend and colleague, also provided us with words of support and encouragement. LAC would like to thank my sons, Miguel and Daniel, for being patient during my summer field seasons and, on occasion, helping in the field. Their constant presence, in body or spirit, gave me the motivation to keep going. My parents, brothers, and sisters have always been there for me, too. Their help, support, and patience with me are admirable and greatly appreciated. I only hope that I at least gave them something to be proud of. LMS would like to thank my husband, Kevin, and my parents and sister for their love and support; without them my trips to the field would not have been possible. I would also

xvi / Acknowledgments like to thank Antonio Curet and Lee Newsom for giving me the opportunity to get back into the field in 2003: little did we know then that I would stay with the project and have the opportunity to contribute on a long-term basis. I would also like to thank the visitors to Tibes, who even on our worst days could reenergize us with their enthusiasm for what we were doing. I would have never thought that so many people would have wanted dirty, stinky archaeologists in their vacation photos!

Tibes

1 Introduction L. Antonio Curet and Lisa M. Stringer

After over a hundred years of research and time served as one of the backwaters of anthropological archaeology in the New World, Caribbean archaeology is experiencing a renaissance. In early times, archaeologists in this region concentrated almost exclusively on the important, but limited, issues of migrations and culture histories. Fairly simple methods of excavation and artifact and data analysis were used, and issues relating to other social and cultural practices were normally ignored or assumed without further confirmation. Since the early 1980s, however, Caribbean archaeologists have begun expanding their investigative horizons by adding a wide variety of research topics, including interaction between human communities (e.g., Crock 2000; Curet 2004; Hofman 1995; Hofman and Hoogland 1999; Rodríguez Ramos and Pagán Jiménez 2006; Torres 2001, 2005), social and cultural processes (e.g., Chanlatte Baik and Narganes Storde 1986; Curet 1992a, 1992b, 1996; Curet and Oliver 1998; Oliver 1998; Siegel 1989, 1991, 1992, 1996, 1999; Veloz Maggiolo 1991, 1993), subsistence systems (e.g., Chanlatte Baik and Narganes Storde 1983; deFrance 1989; deFrance et al. 1996; Newsom 1993; Newsom and Deagan 1994; Newsom and Wing 2004; Wing 2001a, 2001b, 2001c), and social organization (e.g., Boomert 2001; Curet 2002, 2003; Keegan and Maclachlan 1989; Keegan et al. 1998; Tavares María 1996; Valcárcel Rojas 1999, 2002). Furthermore, new field, laboratory, and data-processing methodologies have been applied or developed to ensure the collection of appropriate data to address these topics (see various papers in Hofman et al. 2008). Needless to say, many of the new avenues of research have forced a reevaluation of accepted understandings of the past inhabitants of the region. The present book is about a project that is part of this trend. Particularly, our

2 / L. Antonio Curet and Lisa M. Stringer interest lies in studying changes in social organization in ancient Puerto Rico from a lower level and smaller scale of analysis. As in most regions throughout the world, Caribbean archaeology has traditionally studied changes in the archaeological record using cultures as the main unit of analysis. While useful to develop the general chronology of the region and to define spatial distributions of cultural traits, this approach is not appropriate to study fine-grained processes such as those present in social and cultural changes (see Curet 2003). On the contrary, in order to study many of these processes, research has to focus on smaller scales and lower levels of analysis, in social and cultural units where decisions were made and social and cultural interactions occurred that eventually led to the changes we are studying. This does not mean that we have to ignore higher levels and larger scales, but until recently these units have received most of the attention in the Caribbean, while we know very little about smaller units and the people who composed them. As discussed below, for this reason the Archaeological Project of Tibes has been designed to use households and the settlement as our main unit of analysis. In preparing the research design of the project, two approaches were kept in mind: multidisciplinary and multistage strategies. It is clear that in order to gain a better understanding of the past we need the combined information that can only be provided by other disciplines. Most of the chapters in this volume are the result of this approach, as they are the product of specialists in various fields of study. Although this was the idea since the inception of the project, it has to be admitted that not all these studies were part of the original research design and some of them became affiliated with the project in different ways. The paleoethnobotanical, zoological, soil, and regional studies were planned from the beginning, but other analyses such as the lithology, osteological, bone chemistry, geophysical, and regional studies were initiated by other colleagues. However, these latter colleagues have been working in collaboration with the main project to ensure an efficient way of integrating data from all projects and to benefit each other. Moreover, the evidence collected and the results of the contributions of other colleagues to the project made us change some aspects of our approach. Even though our emphasis was on small units of analysis and lower levels of social groups, it was clear almost from the onset of the project that while many of the processes of interest tended to originate at these levels, they were also integrated within larger units and higher levels (see Torres, this volume). We were forced, therefore, to use a multiscalar approach to the analysis of the data. So, we ask the reader to keep in mind that while most of the chapters included in this volume concentrate on work done at Tibes, the results can be integrated into studies done of other sites in the region to help create a more complete picture. In the rest of this chapter we present the geological, chronological, cultural, and theoretical contexts of the site and the project. The descriptions of these contexts will serve as a backdrop for all of the chapters that follow and serve as a basis for

Introduction / 3 understanding both the issues and the explanations provided in each case. We finish the chapter with a short discussion of the rest of the volume.

Social and Political Change in the Caribbean Although the Caribbean presents many advantages for the study of stratified societies, traditionally this topic has been neglected or poorly studied. With few exceptions, archaeological studies have emphasized cultural history with little attention to prehistoric social organizations. It was not until the 1970s that models for social and cultural changes began to be developed, but even then, the discipline continued to be focused mostly on culture-historical studies. Most of the early suggestions for studying the development of social stratification were environmental and demographic models (Chanlatte Baik and Narganes Storde 1986; López Sotomayor 1975; Stokes 1998; Veloz Maggiolo 1977–1978), in which generally the determinant factors (e.g., population growth and environmental stress) were taken for granted. Nevertheless, correlation between those factors and social and political changes has been little supported in the archaeological record (Curet 1992a, 1993). More recently, several Caribbeanists have taken a more political-economy approach to the study of social stratification and have emphasized sociopolitical, ideological, and economic factors involved in the development of this type of society. Many of these models tend to argue that emerging elite made use of some aspects of religious rituals and symbols to claim a special position in society that, when combined with control and use of economic power (e.g., through feasting), helped them to acquire, consolidate, and maintain political power. However, due to the lack of appropriate information, most of the published arguments have been based on coarse-grained data obtained from ethnohistoric sources, several sites, wide regions, entire islands, or even groups of islands (Curet 1992a, 1996; Curet and Oliver 1998; Keegan 1991; Keegan and Maclachlan 1989; Oliver 1998; Siegel 1996, 1999). Very little effort has gone toward the collection of more refined data at the smaller level of the community or household in order to develop more detailed and realistic models. The approach of using information at the level of culture in the modeling of past human behavior in the Caribbean has multiple deficiencies, four of which are discussed here. First, as is the case in many parts of the world, there may be an overt reliance on ethnohistoric models to describe ancient social and political organizations. In the majority of cases these ethnohistoric reconstructions and models are extrapolated to explain and describe indigenous societies from previous periods and from other regions within the Caribbean. In other words, they become the standard used to evaluate and explain the archaeological record of disparate areas and periods. One correlate of using the early Spanish chronicles to characterize the indigenous groups from the Greater Antilles is the assumption that the in-

4 / L. Antonio Curet and Lisa M. Stringer formation provided in the chronicles is representative of all the societies in these islands. This assumption is based on the premise that Caribbean societies from different islands were culturally and socially uniform with very little diversity. Following this line of thinking, most of the information provided in the chronicles has been applied to the reconstruction of other indigenous societies for which ethnohistoric documents are lacking. An example of this is the use of the terms and concepts of cacique and cacicazgo to describe any “stratified” indigenous society of the Greater Antilles, without addressing the applicability of the analogy. Other scholars have applied this model to the Lesser Antilles (Crock 2000) and Bahamas (Keegan 1992, 1997a), and one extreme application is the use of the cacicazgos of the Greater Antilles as a standard for studying similar societies in other parts of the New World (Redmond and Spencer 1994). Considering that the great majority of the ethnohistorical information was collected from various groups from the island of Hispaniola, it is unclear how much of the cultural and social reconstructions are applicable to other islands, or even to all parts of Hispaniola itself. There are strong reasons to doubt that all polities within Hispaniola and in the rest of the Caribbean were highly stratified and centralized societies (Anderson Córdova 1990; Curet 2003; Tavares María 1996; Wilson 1990). It is argued here, therefore, that although the cacicazgo model is a good starting point for the analysis of past social organizations in the Caribbean, its applicability in different regions and periods has to be tested rather than assumed. Further, in most cases, we have assumed that the only possible form of “hierarchical” societies in the Caribbean is the cacicazgo described by the Spanish chronicles, without considering other possible forms of sociopolitical organization. In recent years, anthropological discussions of nonegalitarian societies have recognized the existence of various forms of social organization (i.e., heterarchy, corporate and network strategies, etc.), including some that contain features from both egalitarian and hierarchical societies as traditionally defined (Blanton et al. 1996; Crumley 1995; Feinman 1995; Feinman et al. 2000; McGuire and Saitta 1996; Saitta 1997). The possibility that some Period III or IV societies were organized in any of these alternate social forms has to be considered. We cannot discard a priori that other forms of hierarchical societies also existed in the Caribbean. Even if all of these groups were organized in cacicazgos as described by the Spaniards, there is evidence indicating that many of them had differential numbers of decisionmaking levels (Tavares María 1996; Wilson 1990). Thus, the political arena of the different islands or even regions within a culture area could have been composed of a spectrum of different types and sizes of middle-range societies instead of “standardized” political units or polities. Also, it is important to recognize that multiple forms of nonegalitarian societies could have existed not only in different islands but also within the same island or region. Another deficiency present in many of the general models suggested for the

Introduction / 5 Greater Antilles resides in the units of analysis used. Most of these models automatically make use of the cultural categories developed by Rouse (1992) for the reconstruction of culture history as the basic social and political unit without a further evaluation. We tend to ignore the reasons for their original creation and assume, consciously or unconsciously, that they are “natural” categories innate to the archaeological assemblages. While the definition of archaeological cultures was, and still is, very useful as a general chronological framework, it cannot be forgotten that they were created for tracking migrations and reconstructing cultural sequences and as such they are not necessarily concepts that should be used for every type of study. Rouse’s categories were developed from a normative perspective that emphasized similarities and differences at higher levels of analysis, i.e., cultures and peoples, levels of analysis that may be inappropriate for the study of social processes that are mostly related to lower levels such as immediate regions, communities, households, or individuals. Competition for status, long-distance exchange, elite interaction, warfare, and other social activities are phenomena that occur at these lower levels and rarely at the level of culture. Many of these categories, moreover, are too encompassing and tend to homogenize significant social, cultural, and chronological variability in the archaeological record that, while not critical for culture history, are of vital importance for the understanding and reconstruction of social processes and organizations. Finally, it has to be recognized that while most of the previous theoretical constructs deal with the “aggrandizers” or social climbers (especially the successful ones), it would be folly to envision the rest of the population as an idle entity shaped and formed according to the will of the power seekers. Since to a certain degree every faction, segment of society, household, and even person has the capacity to make decisions, the rest of the population has the option to resist, react, or conform to the actions of the social climbers. Thus, an important point in the study of social and cultural developments is the overall response (or level of resistance) of other segments of society to the machinations, strategies, and/or changes facilitated by some individuals or groups and how the emerging elite monopolized and sustained various forms of power. Thus, in certain situations (e.g., when power is shared by different institutions), factions or individuals have to negotiate and renegotiate with other segments of society to be able to accomplish specific social goals. This is an issue that has not been addressed by any of the models developed for the Caribbean. Previous studies have highlighted the necessity of developing research strategies flexible enough to assess factors unique to individual case studies, relative to those based on general principles, in order to understand the broader significance of social and cultural developments. On the basis of the points discussed above, it is strongly believed that at least initially this can be accomplished by using domestic groups and communities as our basic units of analysis. It is for this

6 / L. Antonio Curet and Lisa M. Stringer reason that this research project focuses on the study of household and community economy, organization, and composition at the case study site in the Caribbean. Specifically, the project is interested in measuring at the local level changes in household economy, internal organization, and accessibility to economic, religious, and symbolic resources related to the development and internal operation of socially stratified societies. Thus, the study is concerned with determining the role of small groups of individuals in the development and reproduction of institutionalized leadership and how they managed or manipulated natural, economic, political, social, and ideological resources to acquire, increase, and maintain (or share) different forms of power (Blanton et al. 1996; Brumfiel 1992; Cowgill 1993; Earle 1997). At the same time, the project explores the organization of households within communities and the region and how these institutions shifted (or reacted) with changes in the social, political, and economic realms. Another point of interest is determining the environmental, social, and historical conditions that allowed the decisions made by individuals or factions to be successful in controlling different dimensions of social power (Earle 1997). Ultimately, we would like to understand how these small social units are related to the community and society at large by studying how their archaeological correlates relate to the public structures within the ceremonial center of Tibes.

Ancient History of Puerto Rico Although scientific archaeology started in Puerto Rico at the beginning of the twentieth century (e.g., De Hostos 1941; Fewkes 1970), it was not until the mid1930s with the works of Rainey (1940), later expanded by Rouse (1952, 1964, 1982, 1986, 1992), that cultural groups were better defined chronologically, based mainly on ceramic attributes. Rouse (1986, 1992) used a hierarchical taxonomic system in which styles were defined using ceramic modes. These modes correspond to a particular geographic region and chronological period. Styles related in space and/or time are grouped in subseries according to their similarities. Finally, subseries are grouped in series. Since archaeologically Tibes is located in and belongs to the cultural tradition of eastern Puerto Rico, the discussion that follows emphasizes this tradition as defined by Rouse (1964, 1982) and summarized by Rodríguez (1992) and Curet (1992a, 1996; Curet et al. 2004). While Archaic groups had been present in Puerto Rico at least since 3500 b.c. (see Ayes Suárez 1989; Moscoso 1999; Rodríguez 1997, 1999; Rodríguez Ramos 2002a, 2002b), the early Ceramic age is characterized mostly by the Cedrosan Saladoid (300 b.c.–a.d. 600) subseries, which is generally equated with the first horticultural and ceramic-producing groups to migrate to Puerto Rico from the South American continent. In Rouse’s model, the Cedrosan Saladoid in Puerto Rico has three typical styles, Hacienda Grande (300 b.c.–a.d. 400), La Hueca (300 b.c.–

Introduction / 7 a.d. 300), and Cuevas (a.d. 400–600), although there is some debate about the nature of the La Hueca style wherein some scholars are arguing convincingly that it is unrelated to any Saladoid style (see Oliver 1999 for a review of the debate). The Saladoid series is characterized by high-quality ceramics and the use of paint as the main decorative technique. On the basis of the lack of evidence of social stratification in burials and household deposits, most Caribbean researchers consider Saladoid groups to have been relatively egalitarian or tribal in nature (e.g., Boomert 2001; Curet 1996; Curet and Oliver 1998; Keegan 2000; López Sotomayor 1975:103; Moscoso 1986:307; Rouse 1992:33; Siegel 1996, 1999). The Cedrosan Saladoid was followed by the Ostionoid series (a.d. 600–1500). Rather than a foreign migration, this transition seems to be the result of local development in Puerto Rico. Interestingly, by the onset of this series, Puerto Rico had developed two spatially distinct cultural and stylistic divisions, the Elenan and Ostionan subseries (a.d. 600–1200), each with its own styles. The Elenan subseries of the Ostionoid series is associated with both the Monserrate (a.d. 600–900) and the Santa Elena (a.d. 900–1200) styles of eastern Puerto Rico. The Ostionan subseries is divided into the Pure (a.d. 600–900) and Modified (a.d. 900–1200) Ostiones styles (Rouse 1982), which are concentrated on the western side of the island. During the interval of the Elenan subseries, eastern Puerto Rico saw dramatic changes, including a sharp increase in the number of sites, the development of ball courts, plazas, and ceremonial centers (including Tibes), shifts in mortuary practices, and a decrease in the size of houses. The intensity and nature of these changes have led Moscoso (1986:301) and Veloz Maggiolo (1977–1978:59) to argue that they are strongly related to sociopolitical changes, from which institutionalized social stratification emerged. In addition to the general trend in simplification of the pottery assemblage from the Hacienda Grande to the Santa Elena styles, there was also a decrease in aesthetics (in both stylistic complexity and aesthetic priority) and in craftsmanship. Saladoid ceramics tend to have finer paste and are thinner-walled than Elenan Ostionoid ceramics, and the general appearance of the former is more refined relative to the latter. Thus, the tendency in ceramics is of “degradation” of the workmanship and a reduction in the use of symbolic decoration in the form of designs. These gradual but radical changes have been described as a “de-evolution” by Roe (1989) and the later ceramic styles as the “Dark Ages” of the Greater Antilles by Rouse (1982). However, while the pottery is less appealing to the eye, physical analysis suggests that post-Saladoid pottery is better manufactured for utilitarian uses (Curet 1996). The last ancient period (a.d. 1200–1500) in Rouse’s cultural chronology for Puerto Rico consists of the Chican subseries of the Ostionoid series. The Chican Ostionoid is considered to be the archaeological expression of the Taíno groups encountered by Europeans. As in the case of the earlier Ostionoid, this subseries pre-

8 / L. Antonio Curet and Lisa M. Stringer sents regional ceramic variations, in this case at the level of styles and not at the level of the subseries. In western Puerto Rico, the Capá style predominates, while in the eastern area the Esperanza style is dominant. Although both styles are characterized by the use of incisions and combinations of incised lines and punctation, the Capá style tends to have more complex or elaborate designs than the Esperanza style. The Chican Ostionoid subseries continued many of the practices that began in the early part of the Ostionoid series, although on a larger scale and to a greater degree. These included the construction of ball courts, plazas, and ceremonial centers and an increase in the craftsmanship of high-status objects and religious paraphernalia. Thus, in general, artifactual and settlement pattern data suggest that the Chican Ostionoid subseries had more elaborate social and religious systems than earlier groups. This chronological model developed by Rouse has been strongly criticized in recent years (Curet 2005; Rodríguez Ramos 2007). Some of the comments against Rouse’s model concentrate on its unfounded premises and assumptions. However, in recent years Rodríguez Ramos (2007) has presented convincing evidence showing that Puerto Rican chronology was not necessarily unilineal (i.e., having a sequential order of one culture after the other) as Rouse presented but, on the contrary, that on several occasions more than one culture overlapped in time. In other words, cultures did coexist at the same time and region. At the present, the coexistence at Tibes of one or more of the styles described above cannot yet be completely ruled out. However, most of the radiocarbon dates obtained to date for the various ceramic assemblages deviate very little, if anything, from Rouse’s suggested dates. Therefore, throughout this volume we maintain Rouse’s chronology to date units/levels or assemblages for which we do not have a radiocarbon date. Nevertheless, the reader is warned that future dates may significantly alter many of the interpretations of the evidence included here.

The Ceremonial Center of Tibes Geographic Location and Physiology of Tibes Tibes is located near the southern coast of Puerto Rico just north of the modern city of Ponce, approximately 8 km from the coast (Figure 1.1). The site was established on an alluvial terrace off a meander of the Portugués River, and it is limited on the north and east by hills and on the west and south by the river. These limits are physical barriers that protected the site by making accessibility difficult for heavy machinery such as plows. The land where the site is located is known as La Vega del Taní because it is a low, fertile plain produced mostly by the deposition of alluvial soils. Politically, the site is in Barrio (ward) Tibes of the Municipality of Ponce. From a regional perspective, the site is in a biogeographic and geological tran-

Figure 1.1. Map of Puerto Rico showing the location of the Ceremonial Center of Tibes.

10 / L. Antonio Curet and Lisa M. Stringer

Figure 1.2. Topographic map of the region around the Ceremonial Center of Tibes.

sitional zone (Figures 1.1 and 1.2). Geographically, the site is positioned between the Dry Southern Coastal Lowlands and the Semiarid Southern Foothills or piedmont of the Central Mountains (Ewel and Whitmore 1973; Picó 1950). The Dry Southern Coastal Lowlands is a rainfall-deficient area composed of level, fertile alluvial plains with an average annual rainfall around the Ponce area of 917– 955 mm (Picó 1950:94, 1974:370). However, precipitation is unevenly distributed throughout the year, and severe droughts can occur from December to April. The high fertility of these soils seems to be the result of a combination of excellent parent material and an arid climate, which prevents soil leaching or erosion. This geographic region correlates with the Subtropical Dry Forest as identified by Ewel and Whitmore (1973), which is described as the most arid area for the whole island, with deciduous vegetation characterized by thorny shrubs and trees that normally do not exceed 15 m. The Semiarid Southern Foothills lie between the Central Mountains in the north and the Dry Southern Coastal Lowlands. Geologically, these hills include both plutonic and sedimentary formations of late maturity; the former formations are dominated by diorite and the latter by limestone and sandstone. The average width of these foothills along the southern portion of the Central Mountains is 8 km. Although the climate of this region can be considered semiarid, in reality it can be regarded instead as intermediate since it is less dry than the Dry Southern Coastal Lowlands, with an average annual rainfall of about 1,524 mm (Picó 1950,

Introduction / 11 1974). The soils of this region are relatively fertile, but the intermediate aridity limits its agricultural production. This geographic region correlates with the Subtropical Moist Forest as identified by Ewel and Whitmore (1973), characterized by lush vegetation consisting mostly of grasses, scrub bushes, and cacti. However, trees can be as tall as 20 m. The Rainy West Central Mountains (Picó 1950, 1974) consist of rounded landforms of early maturity, with steep slopes, often as much as 40°. They are mostly composed of volcanic and plutonic formations, but hard sedimentary strata are often found, the product of depositions between volcanic episodes. These mountains contain the highest peaks on the island, including the Cerro La Punta at 1,371 m. The temperatures in this region are lower than in the foothills and lowlands, with an average annual temperature of 20.1°C. Rainfall is abundant, with an average annual precipitation of 2,413 mm, which results in a more lush vegetation than in lower elevations. With the exception of some interior valleys and alluvial soils near streams, the soils tend to be shallow and acidic. Although some parts of this geographic region can be placed under the life zone of the Subtropical Moist Forest, most of it can be classified in the Subtropical Wet Forest (Ewel and Whitmore 1973) that occupies the highest areas of the central mountains of Puerto Rico. This zone has high precipitation, with average annual precipitation between 2,000 and 4,000 mm. Vegetation is very dense and diverse, with more than 150 species of trees, some of which are more than 20 m tall. Most of the soils in and around Tibes belong to the Jacaguas and Fraternidad series of Puerto Rico (Abruña et al. 1977; Gierbolini 1979; González Colón 1984:104–105). The Jacaguas series consists of well-drained, nearly level soils on floodplains in semiarid areas very near streams. The soils range from silty clay loam to cobbly clay loam. This series belongs to the Mollisols order that is composed of relatively recent alluvial soils and is considered one of the most productive types of agricultural soils; their reaction ranges from neutral to alkaline, and they are rich in organic matter. The Fraternidad series consists of moderately welldrained soils with little slope that are present mostly on the coastal plain in the semiarid areas of Puerto Rico. The soils are mostly clay formed in fine-textured sediment derived from volcanic and limestone rock. This series belongs to the Vertisols order that includes soil types of recent alluvial soils formed from basic rock material, especially in areas where there is a definite dry season. Vertisols soils, in general, have a high content of expanding lattice clays, but are also naturally fertile. Thus, both types of soils are considered to be appropriate for the production of many crops such as cotton, maize, and manioc (González Colón 1984:105; Roosevelt 1980). Furthermore, the region around Tibes is rich in faunal and botanical resources. Faunal resources in the immediate area around Tibes are dominated by many spe-

12 / L. Antonio Curet and Lisa M. Stringer cies of permanent and migrating birds, but also bats, reptiles, amphibians, and riparian species (fish, crawfish, and land turtles and crabs) are abundant (see González Colón 1984; Maiz López 2002; Newsom and Wing 2004; see deFrance et al., this volume). However, because of the relatively short distance to the coast, Tibes also had access, through direct acquisition or trade, to many marine and littoral resources such as fish, seashells, crabs, sea mammals, and turtles (see deFrance et al., this volume). Botanical resources include a long list of tropical trees with excellent wood that can be used for construction material, canoes, fuel, and medicinal uses (see González Colón 1984; Newsom and Wing 2004; see also Newsom, this volume). In simple terms, Tibes (Figure 1.3) is characterized by being located on a river terrace surrounded by hills and the Portugués River. The general topography slopes gently from north to south. The soils are relatively good for agriculture, but boulders, cobbles, and gravel that occur naturally in some parts of the site make the use of plows or other heavy machinery impractical. Interviews with local people have indicated that the land where the site is located has been used at different times to grow coffee and fruit trees, including oranges, and as pasture for cattle. When discovered, Tibes was covered by thick secondary forest, not much different from the forests that cover some of the surrounding areas today and are composed of native and imported types of trees, including many examples of higuera or gourd trees. Today, the archaeological area has been cleared of the thick forest and a variety of trees native to Puerto Rico were either left in place or new ones planted. Other areas of the property have been left with thick secondary forest or pasture. There are efforts by the current administration of the park to create a botanical garden, where visitors are educated not only on the historical aspects of the site but also on its natural resources. In conclusion, it is clear that Tibes is located in a geologic, ecological, and topographic transitional region in southern Puerto Rico. This prehistoric pattern of settling in transitional or “ecotonal” areas has also been observed for other regions of Puerto Rico, including the late pre-Columbian ceremonial center of Caguana (Carbone 1980; Oliver 1998). It has been suggested that these high-diversity areas were preferred by indigenous people to ensure access to the resources available from two or more distinctive geological, geographical, and/or ecological source areas. However, it is possible also that symbolically these groups may have considered these transitional areas as liminal spaces within the landscape. Thus, these areas may have been regarded as spaces charged with numen or cosmic energy where different parts of the cosmos overlap (i.e., the natural vs. supernatural worlds or the world of the living vs. the world of the dead ancestors). If so, then it is not surprising that in southern Puerto Rico several sites with ball courts and ceremonial centers such as Tibes are located in the foothills (Curet 2005; Curet et al. 2004; Lundberg 1985; Rodríguez 1985; Torres 2001).

Figure 1.3. Topographic map of the Ceremonial Center of Tibes showing the locations and names of the main monumental architecture.

14 / L. Antonio Curet and Lisa M. Stringer

Tibes in the Chronology of Eastern Puerto Rico The trends in the material culture from Tibes follow more or less the general patterns identified for eastern Puerto Rico mentioned above. The ceramic description included here for Tibes is based on the studies conducted by Alvarado Zayas (1981) and Rodríguez Gracia (1998) and a modified version of the modal analysis developed by Curet (1992a). This analysis includes variables and attributes that take into consideration different aspects of decoration, rim form and type, vessel form, temper, and paste characteristics, among others. Additional information on other aspects of the material remains was obtained from these three sources, González Colón (1984), and some of the studies presented in other chapters in this volume. The earliest ceramics discovered in Tibes so far belong to the Hacienda Grande (300 b.c.–a.d. 400) style of the Cedrosan Saladoid series. These ceramics are characterized mainly by white-on-red pottery, composite vessel forms, effigy vessels, and decoration that may include cross-hatching designs. However, due to the limited excavations it is difficult to determine whether these findings come from actual early deposits or they are sporadic finds throughout the site. Independent of the presence of Hacienda Grande deposits in Tibes, it is clear that the presence of the Cedrosan Saladoid subseries at the site is dominated by the later Cuevas style (a.d. 400–600). There is a general tendency toward simplification of the pottery in the transition from the Hacienda Grande to the Cuevas style, leading to a gradual decrease in the use of paints and incisions, and the designs became more geometric. By the end of the Cuevas style, painted decoration was almost absent, although red slips and some complex vessel forms were still common. Unfortunately, it is difficult to assess the beginning and end of these styles at Tibes because no radiocarbon dates are available for any of the Cedrosan Saladoid deposits (see Curet, this volume). The changes during Cuevas times continued into the Monserrate style or the early Elenan Ostionoid subseries (a.d. 600–900). However, the shift from one style to the other is so gradual that by the end of the former and the beginning of the latter, the two styles overlap in stylistic attributes. This has made assigning the assemblages to one style or the other difficult. By this time the white-on-red paint characteristic of the Saladoid series was completely absent and red wares were still present in considerable numbers. Vessel forms were considerably simpler than in the Hacienda Grande and Cuevas styles, most of them being hemispherical in profile, but boat-shaped forms and some animal effigy vessels were also common. Strap and lug handles were still present, although the zoomorphic and anthropomorphic figures seem to have been represented by more abstract figures, which are traditionally identified as heads of bats. In general, the trend in Tibes follows the one observed for eastern Puerto Rico: a tendency toward simplification of the ceramic assemblage with a gradual abandonment of paints and slips leading to the Santa Elena style (a.d. 900–1200) of the Elenan Ostionoid series. In this last style,

Introduction / 15 paints and slips are almost absent from the ceramics. Most of the decoration is in the form of parallel incised lines perpendicular to the vessel rim and “bat head” modeled lugs. As of now, radiocarbon dates obtained from Monserrate and Santa Elena deposits at Tibes fall within Rouse’s chronological designations (see Curet, this volume). Despite the fact that the deposits dated between a.d. 600 and 1200 are dominated by Elenan Ostionoid ceramics, it is important to mention Ostionan Ostionoid ceramics from western Puerto Rico are found throughout the site. The presence of ceramics from one culture area (i.e., western Puerto Rico) in sites from the other culture area (i.e., eastern Puerto Rico) is not uncommon in Puerto Rico and it has been recognized since at least the early work of Rouse (1952). However, little attention has been paid to explaining this phenomenon and determining its significance within the social and cultural history of Puerto Rico. In the case of Tibes, this is more significant, not only because it is a multistructure site but also because it is located in the “gray area” or boundary between both culture areas. This issue is discussed somewhat by Torres in his chapter in this volume. Few ceramics that can be classified as belonging to the Chican Ostionoid subseries (a.d. 1200–1500) have been discovered at Tibes. However, no true deposit dominated by these ceramics has been identified and it is difficult to determine a virtual presence of this subseries. The identification of the presence of this subseries is complicated by the fact that some designs that are traditionally assigned to the Chican Ostionoid subseries (i.e., a.d. 1200–1500) are observable in the late Ostionan Ostionoid subseries (a.d. 900–1200), a phenomenon that José Efraín “Tatito” Irrizarry (personal communication 1990) has called the “Ostiones Achicoidado.” However, the possibility also exists that the site was practically abandoned by this time but was visited regularly by Chican people to conduct religious rituals or other types of ceremonies at this site that remained in the collective memory of many people. Needless to say, more excavation data is needed in order to clarify this issue. Finally, no archaeological deposit belonging to post-Conquest times has been discovered in Tibes. A few dispersed 1800s ceramics, a coin dated to 1864, and pieces of metal have been found either on the surface or a few centimeters below. Also, at least two wells have been found immediately outside the boundaries of the site. Altogether, therefore, and with few exceptions, human activity in historical times seems to have been minimal at Tibes, and most of it was restricted to agricultural practices (see Alvarado Zayas and Curet, this volume, for a discussion on formation processes).

Chapters in This Volume The chapters in this volume come from multiple disciplines and represent in most instances preliminary data or in one case the results of a pilot study to assess the

16 / L. Antonio Curet and Lisa M. Stringer viability of bone chemistry analysis on prehistoric human remains. Some of the chapters also contain information that has never been published. We hope that this preliminary data and analyses will be of help to both professionals and students interested in the prehistory of Puerto Rico and the Caribbean. Chapter 2 provides the reader with a summary of the initial research done at Tibes by the Sociedad Guaynía after the discovery of the site. This chapter is coauthored by one of the first investigators of the site, Pedro Alvarado Zayas, and the current investigator, Antonio Curet. It is hoped that this information from these early investigators, which was previously not widely available, will be valuable to the reader. The history and details of the current project are included in Chapter 3. The discussion by Curet includes an overview of the strategy, the stages of the research, the excavation methods and units, and the integration of the various studies. Tibes was the site of multiple geophysical surveys during two field seasons. Daniel Welch directed a program that used electrical resistivity, magnetic gradiometry, and ground-penetrating radar in an attempt to more accurately and efficiently place excavation units at the site. Chapter 4 gives the preliminary results of these prospection techniques and relates the results of a limited number of excavation units that were placed using the results of these geophysical surveys. The author very honestly relates both the flaws of and success achieved with these techniques and the chapter offers insight for investigators contemplating using these geophysical prospection techniques at sites that have a high concentration of rocks and soil with low moisture content. Lee Newsom, in Chapter 5, presents the results of the analysis of part of the botanical samples obtained by the project between 1995 and 1999. As discussed in Chapter 3, the research design of the current project was developed taking into consideration concerns that the various specialists had. The case of the botanical samples included a very intensive sampling program that is paying off. In Tibes, Newsom has found a wide variety of botanical remains and evidence of behavioral practices in the exploitation of these resources. In terms of wood remains, Newsom was able to identify species used as construction materials, firewood, and fruit trees. Unfortunately, for reasons we do not yet understand, the project has produced very few seeds. Newsom uses this information to understand how the inhabitants of Tibes made use of the natural environment and how they manipulated it to create a new cultural landscape. In Chapter 6, Susan deFrance, Carla Hadden, Michelle LeFebvre, and Geoff DuChemin provide a review of all the faunal material recovered through 2003. The recovered faunal remains have painted a picture of what appears to be foodstuffs used by the group as a whole as well as the occurrence of some items in such small quantities that they may indicate certain vertebrates were reserved for the elite in a social hierarchy or for some ritual or ceremonial purpose. The large quan-

Introduction / 17 tities of marine invertebrate material found at Tibes are also discussed. This concentration of invertebrates is remarkable because Tibes is located approximately 8 km inland. Chapter 7 presents the study of lithics by Jeffery Walker and discusses the analysis of the lithics recovered from 1996 through 1999, including use-wear analysis on some of the pieces to confirm their use. The lithics assemblage at Tibes is not what one would expect from a large ceremonial center. In fact, the quantity of lithics found at the site is very small relative to the overall size of the site. Walker provides a thorough analysis of the lithics and some intriguing suggestions for which activities were taking place at Tibes, as well as indications of possible gender-specific activities. Scott Rice-Snow, Melissa Castor, Andrew Castor, Jeffry Grigsby, and Richard Fluegeman visited Tibes in the summer of 2001 to perform a census of the rocks used to construct the ball courts, plazas, and pavements at the site. Chapter 8 details the results of their study and raises some interesting and as-yet unanswered questions about the sources of the raw material used in some of the petroglyphs found at Tibes. Two chapters deal with the burials at Tibes. In the first of these, Chapter 9, Edwin Crespo-Torres presents the results from a reanalysis of some of the burials found at Tibes during the original excavations between 1975 and 1982. This osteology chapter includes a discussion in which the results obtained from the burials of Tibes are compared with those from burials from other sites. The second chapter dealing with the skeletal remains at Tibes, Chapter 10, presents the results of a pilot study of a larger project conducted by William Pestle of stable isotope analysis. The preliminary results of this study are very exciting, and they raise some interesting questions about the conclusions archaeologists have reached about prehistoric diets based on the faunal remains found at archaeological sites. We look forward to seeing what future analysis of more bone samples will tell us about the paleodiet in Puerto Rico. In Chapter 11, Joshua Torres provides some very interesting information on the interaction sphere for the south-central portion of Puerto Rico. The chapter provides an explanation of the framework Torres uses to examine the relationship between sites in the region. He uses a cost distance model to measure the surface distance between the sites so that he can establish the limits of local spheres of interaction. Torres points out early on in this chapter that people and communities do not exist in a social regional vacuum, which is something that many people tend to lose sight of. Finally, in Chapter 12, Antonio Curet and Joshua Torres attempt to integrate most of the information provided previously in the volume. The discussion revolves around various factors related to the concept of ceremonial centers in the Caribbean, how they are conceived, and what some of the theoretical and epis-

18 / L. Antonio Curet and Lisa M. Stringer temological problems are that need to be resolved in order to understand these phenomena. As mentioned in the chapter, the current project is providing more questions than answers, but the important thing is that it is helping us direct our research questions and designs in the right direction. The chapters in this volume are substantial enough to be viewed as individual research papers, but put together in one place they allow the reader to see that things may not be as they appear. The chapters also show the importance of bringing together researchers with various areas of expertise to provide a more complete picture of the site. Students or those uninitiated with the various specialties represented will find that the authors have provided concise explanations of their specialty so that the reader can understand what is being discussed. Readers who share the authors’ specialty should find that enough detailed information is included to satisfy those who have a greater grasp of the particular specialties. The volume as a whole makes it abundantly clear that there are many questions yet to be answered about the role Tibes played in the prehistory of Puerto Rico.

2 Tibes History and First Archaeological Work Pedro Alvarado Zayas and L. Antonio Curet

The Ceremonial Center of Tibes was first discovered by a resident of the area, Mr. Luis Hernández, who collected dead wood and manufactured charcoal on the land where the site is located. Mr. Hernández became interested in the shells and ceramic “adornos” found on the surface and began collecting them. After doing some bibliographic research he quickly realized that these objects were the material remains left by the pre-Hispanic groups of Puerto Rico (see González Colón 1984:97–98; Rodríguez Gracia 1998:28–30). In August 1975 Hurricane Eloísa hit the island, causing severe floods in the southern part of Puerto Rico, including on the Portugués River. These floods eroded away some of the river silt deposited on the site for centuries and exposed parts of some of the architectural structures. Later that same year, José Efraín “Tatito” Irrizarry, Luis Ortiz Sepúlveda, and Luis Albertorio, members of the avocational organization Sociedad Arqueológica del Sur-Oeste de Puerto Rico (SASOPR), visited the area in their search for new sites. After interviewing several people, they met Mr. Hernández, who led them to the site. The members of SASOPR immediately recognized the deposits, noticed at least three stone structures, and contacted the Sociedad Guaynía de Arqueología e Historia de Ponce, another avocational group. The Sociedad Guaynía visited the site later in 1975 and, under the leadership of Juan González Colón and the supervision of Pedro Alvarado Zayas, began working at the site. This work lasted for seven years and focused on locating, clearing, and restoring the structures and determining the chronology of different parts of the site. The initial research consisted of clearing the area of secondary forest and locating surface structures, but eventually it expanded to include excavations throughout the site. Work at the site consisted primarily of

20 / Pedro Alvarado Zayas and L. Antonio Curet the excavation of 2-×-2-m test units and the clearing and restoration of the stone rows, causeways, and original floors of most of the monumental structures. The results of this work produced new, fascinating, and in some instances controversial evidence, some of which is highlighted in this chapter. Besides conducting the archaeological work at Tibes, the Sociedad Guaynía, with the help of several concerned citizens, began a successful campaign to convince the city of Ponce to create an archaeological park at the site. The Centro Ceremonial Indígena de Tibes officially opened on April 30, 1982, and consists of a museum with permanent and temporary exhibit halls, an archaeological laboratory, a conference room, storage areas, and administrative offices. The services of the park include guided tours of the museum and the archaeological area. An average of 50,000 people from Puerto Rico and abroad visit the center every year, including thousands of school and summer camp children.

Description of the Monumental Structures Although the work by the Sociedad Guaynía discovered 12 stone structures, including ceremonial plazas, bateys or ball courts, and calzadas or causeways, only nine have been restored (see Figure 1.3). The three other structures are one isolated row and two possible fragments of calzadas. Each structure was numbered and some of them were also named. Distinction was made between ball courts and plazas based on the morphology of the structures. In a nutshell, rectangular structures or two parallel rows of stones are considered ball courts because of the nature of the game wherein two sides have to be differentiated (as in the modern games of soccer, volleyball, basketball, and football) and because of descriptions included in the early European chronicles for later groups (see González Colón 1984 and Oliver 1998 for more detailed discussions on this topic; also see RiceSnow et al., this volume). At Tibes there are seven ball courts and they can be sorted into two groups according to their size. Six of these structures (Structures 1, 2, 4, 5, 8, and 9) are relatively small and one is exceptionally large (Structure 3). Quadrangular and rounded structures (Structures 6 and 7, respectively) are considered to be plazas used for communal ceremonies such as dancing and chanting. It is important to clarify that while this typology of structures was created by equating form and function, we cannot dismiss a priori that this relationship is not always true and that different types of activities may have taken place in what we call ball courts and plazas independently of their morphology (see Curet and Torres, this volume, for further discussion on this topic). This section discusses two aspects of the stone structures of Tibes. First, we provide a description of the construction techniques of the stone rows and causeways based on the original excavations and restoration. The second part of the discussion is a brief description of each structure and associated discoveries.

Tibes / 21

Construction Techniques The construction of all of the ball courts and plazas of Tibes involved the leveling of the surface where the structure was being built. In some cases it seems that the leveling was minimal, such as in Structures 1, 4, and 9, while others required the removal of a larger amount of soil depending on the topography of the specific location. The removal of soil in some of the structures was of such magnitude that their interior surface is lower than the surrounding area and they can be considered sunken plazas and ball courts (e.g., Structure 6, Plaza Principal). What makes these structures formal ceremonial or communal spaces are their stone boundaries. There are three types of boundaries: stone rows, causeways, and earthen berms. The most prevalent type present in the structures of Tibes is the stone row. Stone rows are present in almost all of the structures at the site, a practice also commonly used throughout Puerto Rico, the Virgin Islands, and eastern Hispaniola (see Alegría 1983). Judging from the findings during the restoration of most of the structures, it seems that in many cases the first step in the construction of the stone rows was to excavate a small trench in which relatively flat and quasi-rectangular boulders of various sizes were placed standing up on one of their short ends. Smaller rocks were placed as wedges to help level the rocks in order to maintain their vertical position. Later on, the trench was filled with dirt, burying the bottom part of the stones and the wedges. Clearly, the trench and the wedges served as a foundation for the stone row. In some cases, no evidence of the trench was discovered during the excavations, but this could have been because the original surface of the structure was eroded down to the level of the wedges (see Figure 2.1a) or because there was no difference in soil color, composition, or texture between the trench and the surrounding matrix, making it extremely difficult to identify the trench. Causeways are formed of relatively small, flat boulders placed on their flat horizontal surface, producing a paved surface. All examples of causeways in Tibes seem to have been built over a prepared foundation composed mostly of old trash middens. Once the foundation was in place it appears that there were two different methods employed to construct the causeway. The first consisted of placing the boulders horizontally on the ground, shaping the arrangement of the whole causeway to the desired shape (i.e., rectangles or triangles). In the second, two small walls of vertically standing stone rows were first formed in the desired shape. Then, the space between these walls was filled with trash covered with flat cobbles placed horizontally on the surface. The rows of standing stones served both as a retention wall to maintain the integrity of the causeway and as a boundary for the structure. Causeways have been discovered in three of the ball courts (including Structure 3) and in the quadrangular and circular (or star-shaped) plazas. Based on ex-

Figure 2.1. (a) Section of southern stone row of Structure 2 showing wedges and eroded surface; (b) Structure 2 seen from the east; (c) petroglyph below the central section of the western stone row of Structure 6; (d) Unit N276 E105 showing the superficial location of the sprinkler piping.

c

a

d

b

Tibes / 23 cavation data it seems that the construction sequence of all of them began with preparation of the surface by leveling the terrain and preparing a foundation composed of ancient trash or soils mixed with ceramics, lithics, and shells. This foundation was leveled and shaped according to the desired shape and then the boulders were placed to form the paved causeway. In the cases where the causeway was delimited by stone rows, a small trench was excavated to place the stones in vertical position with the help of wedges. These stone rows were probably placed before the foundation of ancient trash. The lithology survey conducted by Rice-Snow et al. (this volume) determined with relatively high probability that, with the exception of one uncommon rock type, the rocks used in these constructions were obtained from the riverbed of the Portugués River or from the site itself. This implies that a labor force was needed to pick up the rocks from the river and transport them to the site under the supervision of someone who knew what type, shape, and size of rocks were needed. Although this may not seem a remarkable deed because of the proximity of the river and the large amount of rocks naturally available, the number and weight of the rocks used in the construction and the specific shape and size needed required a number of man-hours. One case of an earthen berm has been discovered at Tibes, delimiting the western side of Structure 5 (see description below). It is possible that this berm was constructed with the overburden produced by the leveling of the area for the construction of the structure. A final point that needs to be considered in relation to the construction of the structures is their arrangement within the general context of the site and the built landscape. Although it is not discussed here in detail, Oliver (1998) has noted that the arrangement of the structures at Tibes is similar to the one from the later ceremonial center of Caguana. Oliver has argued that this arrangement has some meaning and is related to the indigenous cosmovision and, possibly, the social and political organizations of the groups that used the structures. Based on these similarities he suggests that both centers followed similar canons that were developed in the Elenan Ostionoid subseries at Tibes and then applied at Caguana during the Chican Ostionoid subseries (see Curet and Torres, this volume). Moreover, another study conducted by García Goyco (personal communication 1998) suggests that some of the structures were aligned with astronomical phenomena such as sunrise or sunset in the equinoxes or solstices and, perhaps, the appearance of certain stars or constellations on the horizon. While this is difficult to prove, this combined with Oliver’s arguments strongly suggests that the construction of the structures in the cultural space of Tibes may have involved specialized knowledge of the symbolic use of space, planning, and engineering, and extensive and intensive labor. Although at first the structures at this site seem relatively

24 / Pedro Alvarado Zayas and L. Antonio Curet Table 2.1. General information on the ceremonial structures at the Ceremonial Center of Tibes Structure No. Structure Name 1 2 3 4 5 6 7 8 9

Batey Número 1 Batey de Herradura Batey del Cemí Batey Santa Elena Batey de una Hilera Plaza Principal Plaza de Estrella Batey del Murciélago Batey del Cacique

Length (m)

Width (m)

12.8 35 76 18.4 13 40 30 13.7 15.2

10.9 9.3 15.15 14.1 10.3 37 25 10.4 10.4

Orientation of Long Axis East–West East–West North–South East–West North–South North–South North–South Northwest–Southeast North–South

Note: Information from González Colón 1984. Numbers and names of the structures correspond to those shown on Figure 1.3.

simple when compared to structures from the American Southeast, Mesoamerica, and the Andes, their construction could have involved complex processes and knowledge.

Descriptions of the Structures As mentioned above, the main restored structures from the site were numbered and some of them were named by both Alvarado Zayas (1981) and González Colón (1984) (see Figure 1.3 and Table 2.1). The discussion that follows includes a description of each of these structures, aspects of their restoration, and associated finds discovered during the excavations inside or near them.

Structure 1 or Batey Número 1 Structure 1 consists of a small ball court on the southern end of the site. Its long axis has an east–west orientation. The ball court is located on an old river channel that is still visible on the surface between two old natural levees. The structure is composed of two stone rows on its north and south sides, while its western and eastern sides are open. The north and south stone rows are approximately 12.8 m and 12.65 m long, respectively, and the open sides about 10.9 m. The original surface of the ball court seems to have been about 6 cm lower than the surrounding area. However, it is not clear whether this was intentional or the product of erosion caused by its use and rain. The inside surface of the ball court was devoid of artifacts, but surrounding deposits contained materials belonging to the Elenan Ostionoid period.

Tibes / 25

Structure 2 or Batey de Herradura Batey de Herradura is probably the most complex of the small ball courts. It is located on top of the old river levee and it has a horseshoe shape with only its east side open (Figure 2.1b). The structure measures approximately 9.3 m by 35 m; its long axis follows an east–west orientation. It is bordered on its northern, western, and southwestern sides by a causeway (forming a J or fishhook shape). The east part of the southern side is formed by a stone row. The causeway is delimited on its interior and exterior sides by stone rows and its average width is 1.75 m. A foundation of ceramic sherds, seashells, and rocks (possibly ancient trash) was discovered under the causeway. Rocks were missing from some parts of the interior of the causeway and they were replaced with boulders obtained from the river during the restoration work (see map in Figure 1.3). The interior surface of the structure is lower than the surrounding areas, but it is unclear whether this was produced purposefully or caused by erosion.

Structure 3 or Batey del Cemí Batey del Cemí is located on the southwestern section of the site, in a low part parallel and close to the river (15 m away). Contrary to the two previous structures discussed, Structure 3 is oriented north–south. The eastern and western sides of the structure are partially delimited by causeways; the remaining portions of the east and west sides and the southern side are bounded by stone rows, and the north side is open. The interior dimensions are 76 m long and 15.15 m in average width. Not surprisingly, this ball court was almost completely covered with 50 to 80 cm of alluvial silt. The northern part of the eastern causeway has a triangular extension that reminded some of the original excavators of the shape of three-point idols, hence its name, Batey del Cemí. During the original excavation and restoration of this structure, it was discovered that the western causeway was not at the same level as the rest of the stones in the structure. This causeway was sloping down from south to north, giving the impression that it was built this way purposefully to form a paved trail toward the river. Although this suggestion cannot be discarded, it is also possible that the causeway was originally built at the same level as the rest of the structure but was subsumed later on by mudslides of the river terrace produced by floods. Both causeways were built between rows of small stones and they are placed over a foundation made of ancient trash including shells and ceramics. Test units excavated by the original researchers on each causeway uncovered two human burials. The first burial was found under the northern part of the western causeway and it included two rodent crania and a zoomorphic ceramic head as funerary offerings. The second burial was unearthed from the eastern

26 / Pedro Alvarado Zayas and L. Antonio Curet causeway with the cranium separated from the postcranial skeleton and placed under the thoracic cavity. Excavations were conducted in some parts of the interior surface and 11 human burials were discovered in the middle of the structure in addition to ceramic sherds. These burials were discovered 90 cm below the surface and contained Hacienda Grande–Cuevas ceramics as funerary offerings. Thus, they seem to date prior to the construction of the structure.

Structure 4 or Batey Santa Elena Batey Santa Elena is located south of Structure 2, right on the old river channel. With the exception of the tips of some of the largest rocks, this ball court was completely covered with alluvial silt. Its long axis is oriented east–west. It is composed of two stone rows on the northern and southern sides, while the eastern and western sides are open. The northern stone row measures approximately 14.6 m, the southern one 13.6 m, and they are 18.4 m apart. The interior surface is slightly lower than the surrounding area. Only five original stones of the north side were found, including the ones at both ends; the missing stones were replaced during the restoration work with rocks of similar shape and size obtained from the Portugués River. Test units in each of the open sides of the structure and outside the stone rows produced Santa Elena style material (hence its name, Batey Santa Elena). Earlier ceramics were only found north of the ball court deep under the old river levee just north of this structure, but they do not seem to be related to the construction of the ball court.

Structure 5 or Batey de una Hilera Batey de una Hilera consists of a small depression on the eastern side of the site about 30 m east of the main, central plaza. The long axis runs north–south. The western side of the structure consists of a stone row and the eastern side of a low earthen berm. The interior surface measures 13 m long and 10.3 m wide, and it seems to have been 15 cm below the surrounding area. The north and south sides are open and no cultural material was found on the surface. The berm measured 14 m long, 1.5 m wide, and 8 cm high. Test units in the berm were completely sterile.

Structure 6 or Plaza Principal Plaza Principal is the largest structure in the site and the one with the largest number of rocks. On average it measures 40 m long and 37 m wide, with the longest axis running north–south. The northern and southern boundaries of the structure are delimited by stone rows, while causeways are present on its eastern and western sides. Access to the plaza was possible through openings in the southeastern, southwestern, and northwestern corners; the eastern causeway and the northern stone row are joined in the northeastern corner. Interestingly, trian-

Tibes / 27 gular extensions are observable on both sides of the openings at the end of the causeways. Because of the sloping topography around this structure, a considerable amount of soil had to be removed during its construction. The result of this was a sunken plaza on three of the sides, the western, northern, and eastern sides. The causeways were built with stone rows on their interior and exterior sides to maintain the integrity of the paved surface. Because of the lower interior surface of the plaza, however, the interior stone rows of the causeways were made with larger rocks than the exterior ones. This plaza also includes several petroglyphs, many of them on the interior stone rows of the eastern and western causeways. However, there are a few petroglyphs found on the flat stones that are placed horizontally to form the paved surface of the eastern causeway. With the exception of one possible zoomorphic representation, most of the petroglyphs are simple anthropomorphic faces. Many of these show the outline of the face, eyes, and mouth, and in some cases even ears. Some of them, however, only show three depressions for the eyes and mouth. All but one of the petroglyphs are found on the eastern causeway. The only petroglyph on the western side was carved on a stone partially buried under the middle of the stone row (Figure 2.1c). This petroglyph represents an anthropomorphic face with eyes, mouth, forehead, and ears (or ear spools). It also has lines like rays coming down the chin. Partially buried petroglyphs like this one are rare but not unheard of in Puerto Rico. A classic example is found at the ceremonial center of Caguana at the end of the western stone row of the main plaza. This petroglyph also consists of an anthropomorphic face, but in this case it includes a crown or head decoration, symbolizing the social or ritual status of the individual being represented. Oliver has identified the Caguana petroglyph as Maquetaurie Guayaba, Lord of Coabey, land of the dead ancestors (Oliver 1998:183). The interior of the plaza seems to have been clear of any cultural material. However, excavations conducted west of the plaza produced almost exclusively Santa Elena material, suggesting that this structure may have been built during this time. Interestingly, as in the case of Batey 3, a large cluster of burials belonging to the Cuevas style was discovered under the floor of the plaza. These burials will be discussed later in this chapter.

Structure 7 or Plaza de Estrella Plaza de Estrella is located north of Plaza Principal. Its interior surface is circular or elliptical but is delimited by triangular causeways giving it the shape of an incomplete star or a rising sun. The interior surface is 25 m wide (east–west) and 30 m long (north–south). Evidence from the original excavations suggests that the plaza was originally delimited by eight triangles, but only six of them could be restored. The other two seem to have been destroyed by agricultural practices. The

28 / Pedro Alvarado Zayas and L. Antonio Curet triangles are actually a series of platforms placed over a foundation made of ancient trash fill that includes sherds, seashells, and small stones. Contrary to most of the causeways in the site, the rocks that form these triangular causeways are not delimited by vertical stone rows along their borders. They are simply placed over the surface of the foundation. Interestingly, previous to the construction of the triangles, the area was not leveled; on the contrary, the triangles were built on the inclination of the general topography of the terrain. Two burials were found in the southern part of the structure, under the remnants of one of the missing triangles, north of the northwest entrance of Plaza Principal (Structure 6). This structure was impacted by a bulldozer that was painstakingly brought down the hill north of the site by the owner of the land, before the discovery of the site, to remove rocks for farming the fields. Several of the triangles seem to have been destroyed before the owner discontinued his attempt because of the large number of stones. The original excavators were able to reconstruct one of the triangles by following the foundation of the platform, which retained its original shape.

Structure 8 or Batey del Murciélago Batey del Murciélago is located northeast of Structure 7 and is composed of a stone row on its southern side and a causeway on its northern side. The eastern and western sides are open and the long axis runs northwest–southeast. The north causeway is 13.7 m long, the south stone row is 10.1 m, and the separation between them is about 10.4 m. The stone row is located on a relatively level surface, but the causeway follows the sloping topography produced by the tail of the hill just north of it. The causeway is not straight but curves toward the interior of the ball court. Its construction included the placement of vertical stone rows along the edge of the causeway and it rests over a foundation made of trash midden that included ceramics, small stones, and seashells, including shell artifacts made of Strombus pugilis. One type of shell artifact discovered in this area was several examples of pectorals in the shape of bats with their wings extended, made of the wing of the Strombus gigas. For this reason the structure was named Batey del Murciélago. A human burial was discovered under the southern end of the causeway, 65 cm deep. No cultural material was discovered on the interior surface of the ball court. Only a large stone was discovered in the middle point between the causeway and the stone row, 20 cm deep. The deposits around the ball court, however, produced materials related to the Cuevas, Monserrate, and Santa Elena styles.

Structure 9 or Batey del Cacique Batey del Cacique is located in the northern part of the site. It is rectangular and is formed by western and eastern stone rows, while the southern and northern sides

Tibes / 29 are open. The average size of the structure is 15.2 m long and 10.4 m wide, and its general orientation is north–south. The interior surface is lower than the surrounding area and it is well leveled. Although the immediate surface of the ball court was clean, the structure is located over some of the oldest Saladoid archaeological deposits at the site.

Human Remains The original excavations conducted by the Sociedad Guaynía produced at least 149 sets of human remains. The spatial and chronological distribution of these burials presents interesting patterns that show both continuity and changes in funerary practices. In this section we discuss the archaeological aspects of the burials (i.e., cultural, spatial, and chronological contexts); a detailed report of the biological characteristics and cultural practices reflected by the bones is presented in CrespoTorres (Chapter 9, this volume). To facilitate the discussion, this section is divided according to cultural/chronological periods. However, because of the coarseness of the information used to date these remains, the divisions are at the series level.

Saladoid Burials The great majority of the burials discovered in Tibes belong to the Saladoid series. One interesting pattern observed in the locations of these burials is that they tended to be found in clusters. Two major clusters were identified, both of them under the largest structures of the site, the main plaza, or Structure 6, and Structure 3. The cluster under Structure 6 consisted of 130 individuals, with the highest concentration located on the northwestern corner of the structure. Most of the skeletons were well articulated, with the exceptions of the ones that were found under the western causeway and northern stone row. Furthermore, most of the individuals were placed in some kind of flexed position and many seemed to have been laid over a layer of pebbles (possibly an old river channel). The second cluster, found under Structure 3, consisted of 11 individuals buried in similar fashion to the ones found under Structure 6. Although the number of burials found seems to suggest that the first cluster is larger than the second one, it is important to mention that this difference may have been produced by sampling differences, since many more units were excavated in Structure 6 than in Structure 3. Many of the burials in both clusters included funerary offerings, the great majority of them consisting of ceramic vessels.

Elenan Ostionoid Burials Burials that were assigned to this period were normally found in two contexts: in trash middens or in association with structures. Interestingly, at least three of the burials at Tibes were found under causeways.

30 / Pedro Alvarado Zayas and L. Antonio Curet The funerary patterns observed at Tibes are probably the best example of a generalized trend in mortuary practices identified by Curet and Oliver (1998). Although recognizing that a variety of mortuary practices potentially existed through time and space, these authors noted that in some Cedrosan Saladoid sites in eastern Puerto Rico a relatively large proportion of burials are located in the central clearing of the site. Furthermore, many burials belonging to later periods tend to be located within the household cluster (under house floors or in domestic middens), as seems to have been the case at Tibes. Two unique features of Tibes, however, are (1) the seemingly simultaneous presence of more than one cluster of burials and (2) once the practice of burying the dead in clusters ceased, the superimposition of these areas with ceremonial structures, e.g., stone-lined ball courts, characterized by clearly defined boundaries. Thus, in addition to changes in mortuary practices, there seems to have been a significant shift in the use of space, possibly from a multifunctional space that was used for both profane and sacred activities, to formalized ritual space with clear boundaries that distinguished it from the surrounding areas.

Formational and Transformational Postdepositional Processes One important factor that has to be kept in mind in the study of the archaeological record of complex sites such as Tibes is the role of formation processes. Formation processes are the factors that create the historic and archaeological records (Schiffer 1987). In other words, they are natural and cultural processes that created, shaped, and transformed the sites into the conditions we find them in today. These processes tend to blur patterning in the archaeological record created by the behavior of the people we are studying, making it more difficult to identify and understand them. They can erase any evidence of patterns or even create new patterns, eventually leading archaeologists to wrong conclusions. Although formation processes include the actions conducted by the people who created the archaeological assemblages, in this section we concentrate only on postdepositional processes: natural transformational processes and cultural formation processes.

Natural Transformational Processes Animals, plants, precipitation, and floods are the natural processes that have produced the largest impact on the archaeological record of Tibes. Although Puerto Rico does not have large wild animals, there are many types of insects and invertebrates that may have had an impact on the archaeological deposits at Tibes. The actions of burrowing animals such as earthworms, tarantulas, and other insects, though these animals are small, throughout centuries could have had a major impact by erasing stratigraphic differences or earthen features such as post molds or

Tibes / 31 trash pits, making their identification more difficult. Also, these animals can move artifacts from the subsurface to the surface and the collapsing of their burrows can cause the vertical movement of artifacts throughout the stratigraphic record. Similarly, plants have probably produced changes in the archaeological record of Tibes. For example, tree roots can erase some differences in soils, transpose some artifacts from their original positions, or even create “new” features by leaving marks that look like post molds or pieces of charcoal that are more recent than the archaeological deposits. Moreover, trees uprooted by strong winds or heavy rains can bring to the surface a great number of artifacts, erase stratigraphy, destroy earthen features, and produce depressions that can be confused with trash pits or holes produced by the ancient inhabitants of the site. Possibly the natural process with the strongest impact on Tibes is precipitation. Even under normal circumstances, Tibes is greatly impacted by the effects produced by rain. With moderate rain, the hills north and east of Tibes serve as a basin that collects runoff, directing it south across the site and the structures, eventually draining to the old riverbed on the southern part of the site. In addition to this process, heavier rains can lead to the overflowing of the Portugués River, exacerbating the problem. Two kinds of processes are produced by runoff waters and floods: erosion and deposition of soils. In many cases, the water flows produced by rains are strong enough to induce the removal of soils and in some cases full deposits. Furthermore, upsurge in the flow of water in the river can eventually erode its banks, erase whole stratigraphic sequences, and even destroy some of the structures next to the river. The second impact of precipitation, deposition of soils, is produced by the same processes mentioned in the previous paragraph. Depending on many factors, including amount of precipitation and strength and speed of the water flow, soils and artifacts can be carried, displaced, and eventually deposited by water action. A sterile layer of 15–20 cm of alluvial soil covers most of the western part of the site, indicating that this is a process that has occurred in Tibes (see also Scudder 2001). Also, most of the north and northeastern sections of the site seem to be covered by a sterile layer of colluvial soils that originated from the nearby hills. Interestingly, this erosion and deposition problem does not seem to be new to Tibes and probably was present even during pre-Hispanic times, as evidenced by the apparent eroded surfaces of some of the structures. This deposition and removal of soils affects the archaeological record of Tibes in many ways. The removal of soils may lead to mixing of artifacts from different contexts, collapsing of multiple strata into one, destroying earthen features, removing or mixing artifacts from the deposits, and removing and turning rocks from the ball courts and plazas. It is clear that these processes can greatly impact the archaeological record of Tibes by erasing patterns and evidence and by creat-

32 / Pedro Alvarado Zayas and L. Antonio Curet ing new ones. The identification of the impact of these processes in the archaeological record of Tibes is imperative in order to distinguish real cultural patterns from those produced by natural phenomena. In addition to these main natural processes that could have affected the archaeological record of Tibes, there are also others that are neither visible nor have the same effect as the ones mentioned already. Two examples are chemical and biological processes. For example, it has been determined that the relatively basic (i.e., pH >7) conditions of the soils at Tibes are detrimental for the preservation of some remains such as pollen and phytoliths (Scudder 2001; see also Newsom, this volume). However, these processes are difficult to detect and it is difficult to determine the degree of their effect on the archaeological evidence.

Cultural Formation Processes Cultural formation processes can take a large number of forms. In general, they are less predictable, and in some cases they are less visible, than natural processes. Cultural processes at Tibes may have been produced by the Amerindians who inhabited the site; by post-Columbian, historical populations; and by more recent people. A wide variety of actions by ancient people may have impacted the archaeological record of Tibes. As in any other archaeological site, these processes can range from digging holes for trash pits and postholes, to mining for clay, to leveling floors of houses. However, Tibes has the relatively uncommon impact produced by the construction of multiple ball courts and plazas. As mentioned before, these constructions consisted of leveling off the area by removing and relocating rocks and soils, digging trenches for the stone rows, preparing foundations for the causeways using prehistoric trash, and bringing rocks from the river. The effects of these actions can include the destruction, disturbance, or relocation of previous archaeological deposits, features, and other structures (e.g., post molds) and the burying of previous archaeological deposits. Consequently, the construction of ball courts and plazas could have distorted the archaeological record by “erasing” or relocating older archaeological deposits and features, mixing deposits with different cultural and/or chronological affiliations, using some of these deposits as construction material, and inverting stratigraphic sequences. In postcontact times, it seems that the most common activity at the site was agricultural practices. However, these practices were probably nonintensive (i.e., without mechanical and possibly also animal-driven plows) and low scale because of the large number of rocks present in Tibes’s soils. The extremely low density of historical artifacts at the site seems to suggest a low level of activities during this period and, therefore, less impact to the deposits compared to previous and later times.

Tibes / 33 More recently, before the discovery of the site, one of the owners of the property had an interest in intensifying the agricultural production of the area by plowing it. However, in order to accomplish this, the land had to be cleared of stones and rocks that could break the plow blades. With this in mind, the owner managed to bring a bulldozer through the northern hills. Although it is difficult to gauge the extent of the damage produced by this heavy machinery, it seems that the impact was minimal since almost all the stone structures were spared and since the northern and western deposits are sealed by 10–20 cm of alluvial or colluvial soils. The only known impact was the destruction of one or more of the triangles of Structure 7 on the northern part of the site. It is reported that when the owner observed the amount of boulders and cobbles, he desisted in his intent to clear the land. Despite the destruction caused by the bulldozer, the original excavators were able to reconstruct one of the triangles, because the foundation retained the shape of the causeway. This also suggests that the damage by the machinery was relatively shallow. Finally, other cultural formation processes have been active at the site since its discovery. These include archaeological excavations and activities related to the park. The impact of the original excavations and the restoration of the structures on the integrity of Tibes has been the main topic of conversations among some Puerto Rican archaeologists for decades. While some of these comments may be accurate, many are unfounded, tending to create myths and fallacies. In some cases rumors are created and spread to benefit personal agendas. To clarify many of these misunderstandings, we have decided to discuss in some detail some of the activities that took place during the original excavations and work of restoration at the site. However, it is important that, as the social scientists we are, we take these actions into consideration within their historical context. From the mid-1970s to the early 1980s, professional archaeology in Puerto Rico was practically nonexistent. While many avocational associations existed, the only institutionalized organizations dealing with archaeology on the island were the Centro de Arqueología of the Instituto de Cultura Puertorriqueña and the Centro de Investigaciones Arqueológicas of the University of Puerto Rico, both of them ill equipped, underfunded, and greatly understaffed. Most of the advancements in Puerto Rican archaeology and conservation of sites during this time were conducted, instead, by avocational organizations such as the Sociedad Guaynía, the Sociedad Arqueológica del Sur-Oeste, and the Fundación Arqueológica, Antropológica e Histórica de Puerto Rico. Interestingly, many of the archaeologists working on the island today used to be members of these types of organizations. Most of the archaeology conducted by these individuals at that time was done on weekends and funded with their personal finances. This was clearly the case with the members of the

34 / Pedro Alvarado Zayas and L. Antonio Curet Sociedad Guaynía, who began working at Tibes during their free time and using their own money; later on, when the magnitude and scale of the site were clear, they received support from the Puerto Rican and city of Ponce governments. As with all archaeological projects and because of the nature of archaeological fieldwork, the excavations conducted by the Sociedad Guaynía produced the greatest damage. All archaeological excavations destroy contextual data and impact the matrix where artifacts are present. In addition, they involve the displacement of soils, artifacts smaller than the mesh of the screens used for sifting the soils, and unwanted natural objects such as nonartifact stones. Finally, in many cases excavation includes the addition of modern artifacts to excavated units to signal future excavators that the area has already been excavated or disturbed. This last practice, actually, has helped the current project to determine the precise location of old excavation units. Another activity that affected the site as we see it today has to do with the structures. Beginning in ancient times and continuing until recently, people (preor post-Columbian) may have reused some of the stones from the structures for other purposes or to build new structures. However, the process of the restoration of the structures can also be considered a formation process. During this work, fallen or displaced rocks were reinstated in a location according to the criteria of the archaeologist. On some occasions, empty spaces in causeways and stone rows were filled with river boulders and cobbles, especially in Structures 2, 3, 4, 6, 7, and 8 (e.g., compare map in Figure 1.3 with picture in Figure 2.1b). In the case of Structure 7, a triangle impacted by the bulldozer was restored following the shape of the foundation. Although, in our opinion, many of these actions were well justified within their historical context, we have to admit that they also form part of the processes that shaped the site. In addition to the restoration of the structures, some work that impacted the general landscape of the site was done related to the activities of the park as a public place. Four main works are discussed here: the installation of a sprinkler system, removal of surface stones, the use of heavy machinery, and the filling of the old river channel. In informal conversations, several colleagues have mentioned the installation of a sprinkler system in the archaeological area as a postdepositional transformation process that had a major impact on the site. However, we find these statements a bit exaggerated and based on speculative perceptions instead of solid evidence. The sprinkler system consists of 2.5- to 3-inch (6.35–7.62 cm) PVC pipes and is approximately 1,200 m long on the archaeological site (this estimate does not include public areas surrounding the site). Based on our observations and information provided by maintenance staff, the pipes seem to have been placed in trenches no more than 4 inches (10.16 cm) wide and deep. Actually, these estimates are a little liberal (i.e., overestimated), as many parts of the piping are actu-

Tibes / 35 ally visible on the surface (Figure 2.1d). The concerns that many critics have about the sprinkler system are that the installation of the pipes was (1) too extensive and (2) too deep. These two aspects together imply that a considerable proportion of the deposits were heavily impacted. However, when the length and width of the trenches excavated for the pipes are considered, and using the liberal estimates presented above, the sprinkler system covers an area of at most 120 m2, a very small percentage compared to the 4,350 m2 of the approximate total area for the whole site (less than 2.76 percent). This proportion will be even lower if values of volume are used in the comparison instead of area. Furthermore, most of the sprinkler system is located on the western and northern sections of the site, where the general stratigraphy is superimposed by 15–20 cm of sterile layers of either alluvial or colluvial soils. This means that even though pipes were buried in these areas, they were not deep enough to impact the archaeological deposits. Considering all this, it can be concluded that the actual impact of the sprinkler system on the site is much less than 2 percent of the total volume of the site and, even then, the “damage” was extremely superficial, in most of the cases not even affecting the first stratum. In summary, it is clear that the impact of the sprinkler system is too minimal to affect any major interpretation of the stratigraphic sequence or contextual evidence. The second transformation process related to the activities of the park was the removal of superficial rocks that are distributed over the site. As mentioned at the beginning of this chapter, the soils at Tibes tend to include large numbers of boulders, cobbles, and gravel, and many of them are visible on the surface. This is in part due to its nature as a river terrace and possibly a result of many prehistoric cultural activities, especially the construction of the structures. For public safety and aesthetic reasons, the decision was made to eliminate some of these rocks. Unfortunately, no record was kept on the type of rocks and their location, and, for this reason, it is difficult to evaluate the impact of these actions on the archaeological record. However, the effects to the deposits and the assemblages appear to be minimal. Another transformation process was related to the filling of part of the old river channel on the southern part of the site. As mentioned before, Structures 1 and 4 and the southern part of Structure 3 are located on an old river channel, which occasionally serves as a drainage when the river floods. East of Structures 1 and 4, the channel deepens, forming a small gully. For the safety of the visitors, the decision was made to fill this part of the channel (outside the structures) with alluvial soils. The main impact of this action is mostly on the original topography of the site. The only effect on the deposits, if any, is that a layer of silt has sealed them. Other than that, the impact is minimal. Finally, another major criticism against the original excavations and related to the work conducted to prepare the park for visitors is the use of heavy machinery

36 / Pedro Alvarado Zayas and L. Antonio Curet to level some areas and to aid in the excavations. No heavy machinery was used during the excavation of the site or the restoration of the structures. Machinery was used only outside the archaeological area to level trails for the visitors leading from the museum to the site. Even then, the machinery was used after archaeological tests in these areas showed sterile results. The work consisted of using the front loader of a backhoe tractor to level the trail outside the archaeological area, from the bridge over the river south of the site to just before the beginning of the archaeological deposits south and west of Structure 1. Leveling was also done southeast of the site, in an area where no archaeological evidence has been recovered, for the construction of a model of an indigenous village. Therefore, the site of Tibes has not been impacted by the use of machinery during the excavation process or in the work of preparing the site for the public. Most of the activities were conducted outside the archaeological area or did not impact the deposits or structures in a significant manner. Summarizing, there is no doubt that Tibes, like all archaeological sites with a long history of occupation, has been impacted by both natural and cultural formation and transformation processes since pre-Columbian times. Some of them have been more severe than others. Natural formation processes such as erosion and deposition seem to have been acting on the site since its foundation and continue to affect it to the present. Cultural processes have also been present since the beginning of the occupation and they include the construction of the structures and possibly changes in the general topography of the site. In more modern times, some agricultural activities, clearing of the area, archaeological excavations, and restoration of the structures have had their role in transforming the archaeological record. However, although some of these processes could be considered severe compared to those active at other sites, they do not seem to have affected the assemblages to the point that they are useless for archaeological interpretations. Contrary to what many critics say, the impact of many of these activities is superficial and they have not affected most of the site in a significant manner. More significant activities (e.g., the destruction of the triangles of Structure 7 by a bulldozer) can be easily detected in the archaeological record and considered in the analysis of the data obtained by the original excavators. The discovery of several features and radiocarbon dates obtained by recent research support many of these conclusions (see Curet, this volume).

Conclusions Since its discovery, Tibes has been at the same time a place of fascinating discoveries and one of controversy. Through the course of the excavations conducted in the late 1970s and early 1980s, Tibes debunked some ideas, brought up new discoveries, and created new and enigmatic questions. For example, at the time of the

Tibes / 37 original project it was a widespread “fact” or belief that almost all, if not all, of the ball courts and plazas belonged to the late pre-Columbian period that we denominate as Chican Ostionoid or Taíno culture. While some people were questioning this, most archaeologists tended to support this position. The evidence collected by the Sociedad Guaynía clearly showed that the structures at Tibes were older and it confirmed that even multistructure sites or ceremonial centers were developed at least during the Elenan Ostionoid period. These excavations also discovered the first clusters of Saladoid burials, a pattern that has been reported later for other sites in Puerto Rico. More surprising was to find that these clusters were superimposed by later, ceremonial, structures. These discoveries were the foundations of groundbreaking studies on changes in mortuary practices and community patterns in ancient Puerto Rico such as the ones conducted by Curet and Oliver (1998) and Siegel (1996, 1999). Tibes also raised new questions that forced us to understand that past social behavior was more complex and difficult to understand than what current models made us believe. It forced us to realize that cultural and social changes in ancient Puerto Rico involved concomitant, complex processes that included changes in mortuary practices, continuity in the use of space with changes in its meaning, construction of monumental structures, and changes in ceramic styles. And it raised very important questions such as, Why did this initially Saladoid village develop into a ceremonial center? Why was this important site practically abandoned by approximately a.d. 1200? How does this relate to the widespread abandonment of sites all over the southern coast of Puerto Rico? What were the social, political, economic, cultural, and environmental factors that were involved in its rise and collapse? To answer some of these questions requires different methodology and research designs than the ones used in the original excavations, whose main goal was to identify the structures and the culture history of the site. For questions such as the ones presented above we need more fine-grained data using detailed excavations and more state-of-the-art methods and techniques not available in the 1970s. The purpose of the current project at Tibes is to answer some of these questions using a more refined research design. Although at present we are not in a position to give definite answers to many of these questions, the results of the many types of analysis being conducted have been producing interesting and valuable information about past human behavior at the site of Tibes.

3 The Archaeological Project of the Ceremonial Center of Tibes L. Antonio Curet

The Archaeological Project of the Ceremonial Center of Tibes originated from my interest in studying the development of social stratification in ancient Puerto Rico. In 1992 I concluded a regional study designed to test demographic and ecological models for the development of chiefdoms in the Caribbean (Curet 1992a). Although the study concentrated on a small coastal valley of Puerto Rico, the Valley of Maunabo, it produced results that called into question such models. Instead, I suggested that we should also pay attention to political and social variables in order to understand ancient social changes. However, this kind of study needed a different scale and level of analysis than demographic and ecological models. While the latter depended more on regional data, the former, I argued, needed units of analyses that are more in concordance with social units, such as households and communities. At that juncture I could not think of a better place for studying these variables and changes through time than Tibes. Tibes has a Saladoid, egalitarian, occupation and an Elenan Ostionoid, stratified, occupation in which the construction of ceremonial structures probably began, concomitant with other changes observable in the archaeological record (mortuary practices, house structure, and ceramic styles). At that time it was believed that Tibes was probably one of the best examples of early, emerging chiefdoms not only in Puerto Rico but also in the whole Caribbean. I began communicating with the administration of the park in 1993 and submitted a research proposal to the city of Ponce to begin working in 1994. Because of work being conducted on the collections of the site in 1994, the city asked to postpone the project until one year later. In 1995 Lee Newsom joined the project as

The Archaeological Project of the Ceremonial Center of Tibes / 39 co-director, and she served in that capacity until 2003. In 2004 Lisa Stringer joined the project as lab director. The strategy of the fieldwork presented below was designed by myself and Newsom together. Since then, the project has incorporated a number of specialists and colleagues who have been working with us to various degrees and for various lengths of time. In addition, other scholars conducting independent research at Tibes have worked in partnership with the project. In this chapter I present the history and methodological approach of the project and how the specialists’ research is incorporated into the general research design. This chapter presents most of the fieldwork realized by the project through 2003. It begins with a general overview and short summary of all the work, followed by a more detailed discussion of each of the phases of fieldwork and of the radiocarbon dates obtained to date.

General Overview and Summary of the Project: 1995Ð2003 The research strategy of the project was designed in a multistage format (Figure 3.1). In the first stage, the site was probed to get a better understanding of the general distribution of obvious features (e.g., trash middens) over the site. The stages that followed consisted of searching for and excavating possible household structures and surrounding areas. Probing the site consisted of two parts. The first involved excavating shovel pits every 20 m to detect the presence and position of archaeological deposits (i.e., trash middens) and to determine the boundaries of the site (Curet et al. 2005). The counts and weights of items found in the shovel pits were used to prepare plots of the spatial distributions and relative densities of artifacts and shell across the site (see ceramic concentration map in Figure 3.2). Once the archaeological deposits were defined, a test unit (1 × 1 m) was excavated in each of the midden deposits (Units 1–8) to study them in more detail. A posthole was discovered during the excavation of Unit 8. Because of our interest in studying domestic units, in 1998 we decided to expand this unit. We refer to this excavation area as the Macroblock. Another initiative undertaken during this first phase developed and implemented by Daniel Welch involved the use of geophysical methods to find potential locations of domestic units (see Welch, this volume). We have just started investigating some of the anomalies detected, and the analyses of the excavated materials are still under way. During the 1999 season, Welch conducted a small resistivity study, and some anomalies were detected and examined (Welch 2001). Excavations (Operation 19 [OP19]) in one of these anomalies produced another midden with evidence of what we thought at the time was a hearth. This unit was expanded east and west. In the 2001 season we conducted a more extensive and intensive geophysical project over most of the site. This included ground-penetrating radar, resistivity,

Figure 3.1. Map of the Ceremonial Center of Tibes showing the location of the archaeological excavations conducted by the current project.

The Archaeological Project of the Ceremonial Center of Tibes / 41

Figure 3.2. Ceramic concentration map of Tibes. The contour lines represent 100-g intervals.

and magnetic gradiometry (Curet et al. 2005; see Welch, this volume). In 2003 we began testing some of the anomalies detected by this project. Three new areas were opened up and units were excavated: Unit N276 E105 (2 × 2 m), Unit N215 E70 (2 × 2 m), and Unit N296 E93 (2 × .5 m). We also conducted additional excavations in the original Macroblock and OP19 areas during this season. The first of the new units was located over a Saladoid deposit, while the other two were

42 / L. Antonio Curet positioned near two of the ceremonial structures, i.e., the stone-lined courts or plazas. In the more detailed discussion that follows, units/levels are dated or assigned to a cultural affiliation using mainly the ceramic assemblages. The ceramics were studied using a version of a modal analysis that has 16 variables, including vessel thickness, sherd size, rim type and form, lip form, paste, temper abundance and size, and decoration, among others. Since decorated pottery is found in small quantities, we rely more on the nondecorative variables. Furthermore, although the variables were recorded for each sherd, the final analysis characterizes units/ levels, not individual sherds. In other words, the presence of a few sherds of a given type was generally disregarded when characterizing the provenience (unit/level) as a whole. This was done to avoid putting too much weight on anomalies, since they can be the product of, for example, formation processes; a complete assemblage (if it is primary) should have a more defined or clear, i.e., unambiguous, ceramic “signature.”

Fieldwork: Reconnaissance Phase The aims of the pilot phase, specifically, were to (1) identify the archaeological deposits and the boundaries of the site, (2) determine the nature of these deposits (i.e., domestic, ceremonial, production area, etc.), (3) begin reconstructing the history of the site or community pattern through time, and (4) start detecting and defining the different economic strategies pursued by the prehistoric people at different times. Aims 1 through 3 were necessary to estimate the possible location of the domestic units, while aim 4 was necessary to collect data needed in the reconstruction of prehistoric economies and for the development of models or hypotheses to plan the next phases of the project. The first part of the phase was conducted during the 1995 and 1996 seasons and consisted of probing the site to understand and get to know better the general spatial pattern of the settlement. Probing the site consisted of excavating shovel pits every 20 m. Because of the potential presence of sensitive features and caches, interior areas of ball courts and plazas were not included in this stage. All the soils excavated from the shovel pits were screened using ¼-inch mesh and all artifacts as well as visible faunal remains were collected. Reiterating what was said in the overview above, data obtained from the shovel pits were used to shed some light on the organization of space at Tibes. The weight or counts of the artifactual, faunal, botanical, and other material obtained from the shovel pits were used to draw concentration and distribution maps (Figures 3.2 and 3.3), leading to the identification of at least eight archaeological deposits in the site. Figure 3.2 shows the ceramic concentration map based on the materials recov-

The Archaeological Project of the Ceremonial Center of Tibes / 43 ered from the shovel pits. Two preliminary conclusions can be reached from this map. The first is that at least eight archaeological deposits can be clearly identified. The second is that the arrangement of these deposits seems to form a circle or semicircle around Structures 6 and 7. This concentric pattern has been previously reported for many other Saladoid and Ostionoid sites (Chanlatte Baik and Narganes Storde 1983; Rodríguez 1991; Rouse 1952, 1992; Siegel 1989, 1996; Watters 1994) and is similar to that found for many ethnographic indigenous communities from South America (Heckenberger et al. 2003). This pattern is supported by the distribution maps of lithics, shell, and bone (e.g., see Figure 3.3 for the distribution of shell; not enough lithics and faunal remains were recovered from the shovel pits to produce significant maps). Once the archaeological deposits were defined by subsurface testing, small test units (1 × 1 m) were excavated in each one of the middens during the 1995 and 1996 seasons to determine the nature of the deposits (e.g., domestic vs. ceremonial vs. special activity), their age, and their cultural affiliation(s). In order to maintain good spatial control, each unit was excavated using a combination of natural stratigraphy and arbitrary levels within the deposits (uniform strata thicker than 10 cm were subdivided into arbitrary levels to maintain vertical control). All cultural materials, including artifacts, plant remains, and animal remains, from the excavations were recovered in ¼-inch screens and, with the exception of Unit 1, bulk 10-liter samples of soil from each excavation level (10 cm) for archaeobotanical, archaeofaunal, and soil analyses. Artifacts were analyzed to determine the cultural affiliation and chronology of the different deposits. Most of the data collected during this analysis consisted of diagnostic traits. Due to limitations in time and funding only a small number of the soil samples from selected units were processed. Soil samples were passed through a series of screens to separate the faunal and botanical specimens according to their size. The results of some of these samples show that a wide variety of botanical and faunal species of different sizes were recovered using this procedure (see deFrance et al. and Newsome, this volume). The results of the analysis of the material recovered from the excavation units were used to determine the nature of the deposits, mostly in terms of their cultural and chronological affiliations. Using this information it was concluded that, in terms of the nature of the deposits, deposits A, B, C, E, and F are domestic trash middens while deposits D and G are possible areas where overburden from the construction of the structures was deposited. These last deposits were identified by the large amount of rocks, cobbles, and pebbles mixed with archaeological material. This mixture suggests that previous archaeological deposits might have been destroyed in the construction of the stone-delimited structures. The final midden, deposit H, was identified as the location of a building, based on the presence of a hardened dirt floor and a post mold. Although more studies are needed, our initial study of the recov-

44 / L. Antonio Curet

Figure 3.3. Shell concentration map of Tibes. The contour lines represent 100-g intervals.

ered materials seems to indicate that this structure may not necessarily have been a house (see discussion of the Macroblock below). The results of these excavations were also used to gain some understanding about the general history of the use of space at Tibes. Even though the circular pattern (Figure 3.2) in the arrangement of the archaeological deposits seems neat and clear, we cannot ignore the fact that Tibes is a multicomponent site whose function, use of space, and activities changed through time. This means that this ordering of middens is the final result of the accumulation of materials and formation processes for several hundred years and that they are not necessarily contemporaneous. The data from the materials recovered from the test units were used to determine the nature of the deposits and their chronological position in a preliminary manner.

The Archaeological Project of the Ceremonial Center of Tibes / 45

Figure 3.4. Chronological sequence of the deposits and clusters of burials from Tibes: (a) Late Hacienda Grande and early Cuevas styles; (b) Cuevas and early Monserrate styles; (c) Late Monserrate and Santa Elena styles.

A series of maps (Figure 3.4) presenting the location of the different deposits in each cultural period was prepared to better understand changes in the use of space through time. It is important to emphasize again that while we have detected eight middens, it appears that in the process of the construction of the ball courts and plazas some previous deposits were erased and are not represented in these maps while others were “created” in the same process. These factors could produce a distorted view of the organization of space at Tibes. Most of the materials belonging to the late Hacienda Grande and early Cuevas styles are located in the northern and some of the eastern deposits (i.e., deposits B, C, and E; Figure 3.4a). The diagnostic materials recovered from these periods are few and very fragmentary, especially the Hacienda Grande style sherds. The preliminary results of the ceramic analysis suggest that most of the diagnostic ceramics were present in similar amounts in the three deposits belonging to this period. Cuevas and early Monserrate style materials were collected from the northern, eastern, and western deposits (i.e., deposits A, B, C, E, F, and H; Figure 3.4b), almost forming a circle around the main plaza and the Plaza de Estrella (i.e., Structures 6 and 7). Again, the distribution of the diagnostic ceramics did not show any spatial concentration, most of them being widely spread over the middens dated to these periods. Finally, materials belonging to the late Monserrate and Santa Elena styles were located mostly in the eastern, western, and southern deposits (i.e., deposits A, D, E, F, G, and H; Figure 3.4c). In this case, the preliminary analysis shows possible significant differences in the amount of decorated ceramics in each deposit. So far, it is obvious that the collections from deposit G have a larger number and better

46 / L. Antonio Curet quality of decorated ceramics. This differential distribution of decorated ceramics might indicate the presence of social differences between the households that produced the different deposits or differences in the function of space (e.g., ritual vs. domestic). The closeness of deposit G to Bateys 2 and 4 and the identification of this deposit as an overburden deposit suggest that this is a secondary deposit. Thus, if this deposit can be considered primary, it probably was impacted by the construction of the structures. Contrarily, this deposit may represent the remains of ritual or communal ceremonies. It is interesting to note, however, that Structure 2 is the second most elaborate ball court (after Structure 3; Structures 6 and 7 are considered plazas) and the most complex of the smaller ball courts.

Fieldwork: The Macroblock As mentioned above, the Macroblock excavation area was part of an expansion of a 1-×-1-m test unit, number 8 in deposit H (Figure 3.1). Since the discovery of a post mold and a potential floor in this unit suggested the location of part of a possible domestic structure, we concentrated part of our efforts on this area during the summers of 1998, 1999, and 2003. This part of our second stage of operations at the site was specifically intended to build a finer-grained resolution and understanding of individual domestic deposits. The excavation area is located a few meters north of Structure 3 and west of Structure 6. To initiate the excavation of this broader area, we first reopened the original test unit, and its profiles were then used as guides to direct the process of extending the excavation area. The stratigraphy of the original unit consisted primarily of three strata: (1) a relatively culturally sterile stratum composed of light brown alluvial sediment deposited by consecutive river flood events, spanning from ground level to an average of approximately 20 cm below surface; (2) a much darker cultural stratum containing low densities of a fairly wide range of archaeological remains spanning from approximately 20 cm to 30 cm below surface; and (3) a culturally sterile stratum, very similar in color and texture to stratum one, starting at about 30 cm below surface. While culturally sterile, post molds and possible pits intruded into this third stratum from the cultural stratum above it. The ensuing strategy of excavation was designed to make the fieldwork more efficient and focused very much on the broad stratigraphy. Eight 1-×-1-m units were excavated around the original test unit, forming a 3-m2 area (Figure 3.5a). An additional 1-×-1-m unit was excavated 1 m north of the main excavation block. We maintained stratigraphic control according to the “natural” levels, but used arbitrary levels within these broader strata to tighten the resolution of the deposits. So as to carefully control the excavations without sacrificing time, we removed the upper essentially sterile layer as quickly as possible to devote more time to carefully excavating the floor and related features, as well as deposits immediately su-

The Archaeological Project of the Ceremonial Center of Tibes / 47

Figure 3.5. Sketch of excavation units: (a) the Macroblock; (b) OP19.

perior and inferior to the floor level proper. The upper bulk stratum consisted almost exclusively of culturally sterile alluvium and was excavated to a depth of 18 cm below surface, or just above the culture-bearing deposit. This first stratum was removed from all of the eight units surrounding the original test unit, as well as the unit to the north. The second level was considered to be the immediate level

48 / L. Antonio Curet above the original floor and consisted of approximately 2 cm that remained of the upper alluvial stratum before reaching the floor or second stratum. The third level consisted of the surface deposits of the floor and included the first 2 cm of that stratum. The fourth excavation level consisted of the rest of the floor, in other words, from the end of the last level to the beginning of the sterile layer (from about 22 cm to 30 cm). All units were excavated using hand tools. Railroad picks, trowels, and geology picks were used to remove the first level. Geology picks, trowels, and brushes were used in the last three levels. The soils from the first level were screened using only ¼-inch screens. For the deeper levels, a ¹⁄8 -inch screen was used in combination with the coarser screen. All artifacts and ecofacts were collected from the screens. In addition, 5- to 20-liter sediment samples were collected from each level/unit for specialist analyses, in keeping with the multidisciplinary objectives of the research. In general, the excavation of this area was very slow in comparison with other projects in other sites. There are three reasons for this. The first is that the general strategy of excavation is to excavate in a meticulous manner in an attempt to detect the hard-dirt floor, post molds or holes, and artifacts resting directly on the floor. Second, because of the nature of the soils and the formation processes of Tibes, the detection of compacted dirt floors and post molds is very difficult. Compacted dirt floors seem to have been softened by the action of earthworms, insects, spiders, and roots throughout the centuries. Also, because of the color and texture of soils, postholes and post molds are not clearly observable on the floor plan, even in the lighter sterile layer; on many occasions they needed to be sprayed with water in order to make them more visible. Third, the soils in Tibes are mostly clayey silt or loam, a type of soil difficult to excavate even with railroad picks. These factors made the excavation of house locations at Tibes slower than in other sites. Although these excavations are still unfinished, in general, the units around the possible floor and post mold discovered in 1996 produced a relatively low concentration of artifacts. This is highly consistent with results of other studies focused on activity-related deposits, including floor or traffic areas (Nielsen 1991; Stahl and Zeidler 1990). Moreover, several sherds were found standing on their sides and the cultural stratum had a large amount of small pebbles and gravel. At least four or five additional postholes and post molds were found in several of the units, including a fragment of carbonized post (see Newsom, this volume). All of this suggests that this area may have been the location of a structure used for some type of activity, but considering the amount of refuse it was not necessarily a house. Perhaps it was a roof to provide shade for working outdoors. Contrary to what we discovered in the Macroblock, in houses, relatively larger materials tend to get pushed aside or removed from the immediate area, which is often charac-

The Archaeological Project of the Ceremonial Center of Tibes / 49 terized by compacted soils in which only very fine materials remain. Most of the artifacts encountered during our excavations were small, and very few of them were larger than 4 cm in length. Interestingly, however, this area produced several stone artifacts, including a number of flakes, scrapers, and one small ground-stone axe (see Walker, this volume). In contrast, natural remains, including unmodified stones, small fragments of carbonized wood, and generally fragmentary faunal materials, were relatively frequent. River cobbles and small pebbles were the most abundant items and they were widely distributed over the area, occasionally in apparent concentrations. Some of the stones were fire-cracked (i.e., they showed discoloration and fracture patterns typical of exposure to high temperatures). Charcoal occurred widely but appeared in greater concentration in the western units (see Newsom, this volume), while shells and larger, especially whole, animal bones were relatively scarce. Ceramics were predominantly Santa Elena style (a.d. 900– 1200) and uncalibrated 14C dates on samples range from 880 to 1080 b.p. In general, the small size of the artifacts and the finely crushed or fragmented nature of the charcoal, vertebrate remains, and invertebrate remains are also consistent with our interpretation of this deposit as representing an activity floor or surface. Furthermore, the presence of several sherds standing on their sides and a large amount of small pebbles and gravel suggests that the cultural stratum was part of an earth project in which old trash middens and soils from other areas may have been used to artificially level the area. While originally it was believed that this area was the location of a domestic unit, the data at hand do not support or disprove this belief. Therefore, until more evidence is obtained, we cannot discard other possibilities such as the presence of structures used either in ceremonies or during the festivities that took place at Tibes. The possibility also exists that the postholes discovered in this location belonged to a structure dating earlier than the ceremonial structures or the transformation of Tibes into a ceremonial center (see discussion below).

Fieldwork: Geophysical Survey As mentioned above, to supplement the discovery strategy of shovel pits, soil resistivity was used in the 1998 field season to probe large portions of the site to detect other possible household/domestic areas (see Welch, this volume, for more detail on this project). Although traditional methods were successful in locating deposits and recovering materials to characterize them, they tended to be time consuming, expensive, and in many cases unsuccessful as discovery techniques. The geophysical study was initiated, directed, and supervised by Daniel Welch (2001; see Welch, this volume), at that time a graduate student from Boston University, and covered a total of 5,000 m2 of the site. The study provided a high-resolution “map” of subsurface anomalies (Curet et al. 2003; Welch 2001). Initially, test units of

50 / L. Antonio Curet .5 × 2 m were excavated to investigate extremely strong signals or “hot spots” detected by this technique. However, it was obvious after the first excavation that the hot spots were produced by large boulders or piles of rocks buried under the surface. Although this provided interesting information and expanded our understanding of the internal organization of the site, it did little to further our primary objective of identifying additional domestic units and areas of high activity. Thus, we turned to shovel pits to investigate more areas, including those showing extremely high signals. Test signals of lower intensity were still investigated with test units (Operation 19; see description of the excavation below). In general, this initial geophysical survey was successful in locating some significant areas very quickly, including domestic trash middens, a trail, and a cooking area. Based on these results a more comprehensive and intensive geophysical program was designed and conducted over almost the totality of the site in the summer of 2001. Three techniques that measured different physical properties were used by the project: ground-penetrating radar, electrical resistivity, and magnetic gradiometry. Although some small initial excavations were conducted in 2001, it was not until the summer of 2003 that a concerted effort of excavations was conducted to investigate some of the results of the geophysical study. Three units were excavated: Unit N276 E105 (2 × 2 m), Unit N215 E70 (2 × 2 m), and Unit N296 E93 (2 × .5 m) (see description of the excavations below). Not surprisingly, the three techniques used produced mixed results. In some instances, two or three of the techniques detected the same anomalies, while in other instances a signal was obtained from only one of the techniques. Obviously, this is not an indication of the inappropriateness of the techniques but a function of the disparity of the physical properties that they measure. Therefore, the application of all three techniques was useful in that some anomalies were identified by more than one of them and also because they complemented each other in detecting features that were not visible by all of them (see Welch, this volume, for a more detailed discussion on methodology, results, analysis, and interpretations).

Fieldwork: Excavations of Geophysical Anomalies Several of the anomalies detected by the geophysical program were tested with excavations. This section discusses the most prominent and the ones that have produced relevant cultural and social data about Tibes.

Operation 19 One particular trend observed from resistivity signals of the 1998 season that we came to describe as “medium” intensity was a linear pattern that began close to the edge of the riverbank on the southeast side of the site and headed northwest to an area with a small, circular medium signal. Welch identified this linear pattern as

The Archaeological Project of the Ceremonial Center of Tibes / 51 a possible trail heading from the river to some sort of feature in the site. Two .5-×2-m test units were excavated to investigate the nature of the features, one crosscutting the linear pattern (OP19C) and the other intersecting the circular signal (OP19B) (Figure 3.5b). Both units are located west of the northwestern entrance of Structure 6. While more extensive excavations are needed, the first excavation seems to have confirmed the presence of a trail or path detected by the resistometer, having revealed the presence of a narrow, linear, compacted stratum that is in fact very pathlike. The second excavation, however, exposing part of the circular signal, uncovered what at that time we identified as a possible portion of a hearth. Because of our interest in domestic units and the strong suggestion of the presence of a cooking area in or near Unit OP19B, we expanded the excavation with four 1-×-1-m units; two located east and two west of the original excavation trench. These excavations began in 1999 and continued until 2003. The profiles of the original excavation were used as guides to direct the process of extending the excavation area. All cultural materials from the excavations were again recovered using ¼- and ¹⁄8 -inch screens, along with bulk 5-liter samples of soil from each excavation level (10 cm) for archaeobotanical and microfaunal analyses. Aside from the 5-liter samples, botanical (i.e., charcoal) and faunal specimens were again collected separately, either in the excavation screen or directly in situ. The stratigraphy of the general area consists thus far primarily of two strata: (1) a relatively culturally sterile stratum composed of light brown alluvial sediment deposited by consecutive river flood events, spanning from ground level to an average of approximately 20 cm below surface and (2) a much darker cultural stratum containing high densities of a fairly wide range of archaeological remains and beginning approximately at 20 cm below surface. Because of the richness of the deposits we have been unable to reach sterile soil, so it is not clear how deep the archaeological deposit is. The strategy of excavation used here was somewhat similar to the one used in the Macroblock. We maintained stratigraphic control according to the “natural” levels but used arbitrary levels within these broader strata to tighten the resolution of the deposits. So as to carefully control the excavations without sacrificing time, we removed the upper essentially sterile layer and excavated to a depth where artifacts from the culture-bearing deposit began to appear. Interestingly, this initial sterile layer reached a depth of about 20 cm in the western units and about 30 cm in the eastern ones. The rest of the units were excavated using the combination of natural stratigraphy and artificial levels within the deposits (uniform strata thicker than 10 cm were subdivided into arbitrary levels to maintain spatial control). The soils excavated from these excavation units were passed through ¼- and ¹⁄8 -inch mesh screens for the collection of cultural and natural specimens and artifacts. All material culture from the excavations, including artifacts and botanical and faunal remains, were recovered and 10 liters of soil were collected from each

52 / L. Antonio Curet excavation level (10 cm) for botanical and faunal analysis. Large botanical (i.e., charcoal) and faunal specimens were also collected separately. Although initially it was thought that the OP19B unit included a hearth, further excavations seemed to suggest that the original unit was instead placed in the area between the floor of an activity area, represented by a relatively clean surface, and a refuse midden, characterized by highly dense deposits. This explains why the upper sterile layer was deeper in the eastern units (living floor) than in the western ones (trash midden). Pieces of a large buren or cassava griddle were discovered on the floor, suggesting de facto remains of a cooking area. The midden contained relatively high densities of artifacts, including cooking wares, as well as charcoal, ash concentrations, fire-cracked rocks, animal bones, and shells of edible species, some of them clearly burnt. Further, some areas of the living floor showed orange colorations produced from being exposed to high temperatures. This suggests the presence of a hearth nearby. The density of food refuse (including burnt bone, shell, and seeds; see deFrance et al. and Newsom, this volume) and the presence of the cassava griddle and the abundant fuel wood remains seem to indicate that cooking was the main activity performed in this area or nearby.

Unit N276 E105 The purpose of Unit N276 E105 was to investigate an anomaly detected by the geophysical measurements collected in 2001; it is located southeast of Structure 9 and north of Structure 7. Excavation procedures were similar to the ones used for the 1-×-1-m test units. The unit measures 2 × 2 m and three strata were identified: (1) from surface to 30 cm, a layer composed of dark brown/gray colluvial silt, washed down from the northern hills; (2) from 30 cm to 70 cm, a cultural layer of dark brown clay; and (3) starting at 70 cm below surface, large concentrations of pebbles, cobbles, and boulders, possibly representing an old riverbed. Based on ceramic analysis, this deposit seems to represent an early Saladoid component in the site. The ceramics included examples of both the Hacienda Grande and Cuevas styles. Large amounts of stone artifacts, ceramics, and faunal remains were unearthed from this unit, suggesting that it is a domestic deposit.

Unit N215 E70 Unit N215 E70, measuring 2 × 2 m, also was placed on an anomaly identified in 2001 by the geophysical program. It is located just north of the northwestern entrance of Structure 6. The excavation strategy used in this unit was similar to the one followed in the previous unit. In general, three strata were identified: (1) from surface to 50 cm, a grayish-brown, compacted, medium-fine clayey silt; (2) from 50 cm to 125 cm, a dark brown, fine clayey silt (including a trash pit); and (3) from 65 cm on, a light brown, yellowish sandy-gravely soil. The first stratum

The Archaeological Project of the Ceremonial Center of Tibes / 53 seems to be similar to the alluvial silt deposited by the river on the western side of the site. However, various artifacts were found in low concentrations, including two large pieces of ceramic vessels. The second stratum seems to be the cultural layer. This stratum was composed mostly of two features that intruded into the third stratum (see below). The soil of the second stratum was dark brown, similar to other cultural layers found in the site, but in this case the concentration of artifacts was low. The third stratum, with sandy soils, contained large concentrations of gravel, pebbles, cobbles, and boulders, suggesting that it was deposited by the river and that it may represent an old riverbed. Because of the low artifact density in the cultural layer, it was difficult to identify a clear archaeological deposit. However, excavations in 2007 unearthed two clear features, possibly belonging to different periods. The first feature was the skull of a burial that penetrated the northwestern wall of the unit. These remains were discovered at a relatively shallow depth of 62 cm below surface. Because of limitations in time, we were able to remove only the skull; excavation of the rest of the body is planned for a future field season. A preliminary analysis conducted by William Pestle and Edwin Crespo-Torres suggests that the skull belongs to a 35- to 59-year-old female individual. The presence of cranial deformation and its proximity to the cluster of burials discovered in the 1980s under the main plaza suggest it belongs to the early part of the occupation or during the Saladoid occupation. The second feature consisted of a large conical pit that covered most of the unit in the upper levels and decreased in size to a depth of 143 cm below surface. The pit did not contain a high concentration of cultural remains but rather cobbles, pebbles, and boulders, some of them too heavy to be removed by the team of excavators. Archaeological material found in the pit included low densities of ceramics and lithics, shells (especially Turritella variegata), faunal remains, and relatively large quantities of charcoal. Several decorated sherds and half of a small “effigy” vessel were discovered. A fragment of a possible shell trumpet was also found. This material may have been related to rituals, ceremonies, or events that took place in the plaza. Decorations on some of the ceramics seem to have Capá style (Chican Ostionoid) designs (a.d. 1200–1500), but they may represent terminal Santa Elena/Modified Ostiones styles (a.d. 900–1200). A sample from level 4, toward the top part of the cultural layer (second stratum), where many of the decorated ceramics were found, was 14C dated to 750 ± 40 b.p. (uncalibrated; see discussion on radiocarbon dates below). Traditionally, it has been assumed that Tibes was practically abandoned by a.d. 1200. If true, then this may be one of the latest dates obtained from the site. There is a possibility that even though the site was abandoned, later groups in the area may have visited it to conduct some ceremonies. The nature of the pit, however, has not yet been defined.

54 / L. Antonio Curet

Unit N296 E93 Unit N296 E93, measuring .5 × 2 m, was located in the middle of the southern entrance of Structure 9 to investigate an anomaly detected by the geophysical program. Similar signals were also detected in other ball courts. The excavation strategy was consistent with the one from the previous two excavation units. Five strata were identified in this unit: (1) from surface to 30 cm deep, a dark graybrown, loose clayey, colluvial soil; (2) from 30 cm to 55 cm below surface, a dark gray clayey, colluvial soil, similar to the one above but with some sedimentary inclusions and higher concentrations of archaeological materials; (3) from 55 cm to 80 cm below surface, a lighter brown soil, still containing sedimentary inclusions; (4) from 80 cm to 100 cm below surface, a siltier and lighter brown soil with pebbles, cobbles, and boulders; and (5) from 100 cm to 110 cm below surface, light brown silt with boulders and cobbles. The first stratum was composed of loosely compacted clay. Very few artifacts were recovered from this stratum. The lack of compaction of this stratum may have been related to the excavations undertaken in the restoration of the ball court. Stratum two had a color similar to the first one, but it contained sedimentary inclusions and a higher concentration of archaeological materials. The soils from these two strata seemed to be colluvial in nature, washed down from the hills north and northeast of the site. The sedimentary inclusions were of the same color and consistency as the soils from the hill north of the site. The lighter color of the third stratum seemed to be produced by an increase in the silt content; however, the main matrix seemed to still be colluvial clay. This stratum was archaeologically sterile. Stratum four was lighter in color and included river boulders that continued to the next stratum. The last stratum was light brown silt with boulders and cobbles. This stratum was also sterile. The excavations produced two river boulders placed in the first and second strata, next to each other, in a Saladoid deposit underlying the ball court (see illustration in Welch, this volume). While the shape of the first boulder seemed to be natural, the second one showed what we thought was evidence of human modification. This second boulder had a relatively flat shape with one face seeming to be natural, but the other face was covered with depressions and three of its sides were relatively flat. At first it was thought that the depressions were evidence of grinding, but the roughness and pitted nature of the surface did not fully support this possibility. While at first look the depressions and flat surfaces suggested that this boulder was worked by humans, the distribution of patina and erosion indicates that all these features may be natural (Reniel Rodríguez, personal communication 2006). The actual function of these boulders is still unknown. Interestingly, a concentration of animal bones was found under one of the rocks, but it is not clear whether this was an intentional deposition by the inhabitants of the site or the re-

The Archaeological Project of the Ceremonial Center of Tibes / 55 sult of natural formation processes, wherein the boulder protected these remains from being washed out by water flows. Although the shape of these rocks may be natural, their location (at the entrance of a ball court) and nature (river boulders) and the presence of other artifacts tend to suggest that this feature was a cache that was buried purposefully in this particular place. The ball court most probably belongs to the Elenan subseries, but the deposit in this unit produced only Saladoid (Cuevas) materials.

Radiocarbon Dates We have obtained 11 radiocarbon dates (all calibrations were performed by Beta Analytic, Inc.) from charcoal from different test units and levels at Tibes (Table 3.1, Figure 3.6). Two of the dates come from different levels of deposit A, one from deposit C, six from deposit H (one from a post mold [96-1], one from a charred fragment of a post [03-Feat. 2], and four from the cultural layer), one from Unit N215 E70 (03-1744b), and one from a cooking midden of OP19E (99-5). The dates include a calibrated date of a.d. 770 obtained from a Monserrate style (a.d. 600–900) deposit and a late calibrated date of a.d. 1270 associated with late Elenan or early Chican pottery (ca. a.d. 1200). Eight other calibrated intercept dates fall very close to each other at ca. a.d. 1000–1200. One of these last dates (Table 3.1, No. 97-1) comes from a mixed level that contains mostly late Saladoid (i.e., a.d. 400–600) materials but also later Elenan Ostionoid (a.d. 600–1200) ceramics, strongly suggesting that this deposit is related to a secondary midden that may have been relocated around this time. Two dates were related to structural elements (i.e., one from a posthole and the other from post wood) in deposit H that were located spatially close to two of the stone structures (Structures 3 and 6). The posthole (Table 3.1, No. 96-1) was filled with dark soils containing a mix of some material culture items and charcoal, giving the impression that the post was removed and the hole filled with dirt and trash. The charcoal used for dating the posthole was obtained from this filling. The post wood (Table 3.1, No. 03-Feat. 2) was part of a cylindrical fragment of the bottom part of a charred post. The clustering of dates around a.d. 1000–1200 suggests that a major phase of spatial rearrangement at the site may have occurred during this interval or period of time. The fact that at least one of the dates came from a secondary (earlier, redeposited) deposit and another from trash accumulated in a posthole suggests that these activities may have included removal of older buildings, the construction of stone structures and possibly new buildings made of perishable materials, and the mixing and redistribution of existing deposits (Curet et al. 2006). If this assessment is correct, the agreement in these chronometric dates across the site implies that this spatial rearrangement occurred in a short period of time and that it was

Description

Unit 8, post mold Unit 3, level 5 Unit 1, level 6 Unit 1, level 3 N93.95/E98.05, level 3 N93.95/E98.05, level 4 N94.05/E98.05, level 3 N94.05/E98.05, level 4 OP19E, Feat. 5, level 3 N215 E70, level 4 N184 E55, level 6, Feat. 03-2

Ref. No.

96-1 97-1 97-2 97-3 99-1 99-2 99-3 99-4 99-5 03-1744b 03-Feat. 2

H C A A H H H H N/A N/A H

Deposit 103329 109679 109680 110631 136324 136325 136326 136327 136328 198876 198877

Beta Analytic No. 880 ± 50 890 ± 40 1270 ± 40 900 ± 60 950 ± 40 1040 ± 50 1080 ± 60 1010 ± 40 930 ± 40 750 ± 40 990 ± 40

Radiocarbon Age (years b.p.)

Table 3.1. Radiocarbon dates obtained from four different contexts at Tibes

1180 1175 770 1170 1040 1050 980 1015 1050, 1095, 1140 1270 1020

Calibrated Intercept (a.d.)

1030–1265 1035–1245 670–875 1015–1265 1005–1085 895–1040 855–1035 980–1040 1015–1205 1220–1300 990–1160

Calibrated Date (2d , a.d.)

Figure 3.6. Graphic distribution of calibrated radiocarbon dates for Tibes.

58 / L. Antonio Curet accomplished within 100–150 years or less before the abandonment of the site. If true, these data suggest the site experienced a very rapid growth and subsequent collapse of the sociopolitical system.

Conclusions As mentioned at the beginning of this chapter, the main goal of the Archaeological Project of the Ceremonial Center of Tibes is to study social and cultural changes in Puerto Rico, especially the development of social stratification. More specifically, in the case of Tibes this involves the study of the transition from a village-type settlement to a ceremonial center with multiple stone structures. To accomplish this goal it was decided to use domestic units as our unit of analysis and to develop a multistage and multidisciplinary research design. This research design also acknowledged from early on that detailed data were needed in order to gather the information necessary to answer some of our questions. The recovery of the necessary data would require meticulous field strategy with strong and fine spatial and chronological controls. Furthermore, a series of specialized analyses were needed to complement the field data. Thus, in addition to the fieldwork, the research design of the project included a series of specialized analyses conducted by a number of researchers. Some of these specialists were recruited by the project or came as a consequence of some collaboration with the project: Lee Newsom, paleoethnobotanical analysis; Susan deFrance, faunal analysis; Daniel Welch, geophysical techniques; Joshua Torres, regional and geographical information system (GIS) studies; Jeffery Walker, lithic analysis; and Sylvia Scudder, soil analysis. With the exception of Scudder, all these specialists present some of the results of their studies in the following chapters of this book. Scudder has already published the results of her study in the Caribbean Journal of Science (Scudder 2001). In addition to these scholars, other researchers have been conducting studies parallel to and in partnership with the project. The first is a group of geologists from Ball State University in Indiana who conducted a lithology analysis of the boulders used in the construction of the structures at Tibes. This project was an idea of Scott Rice-Snow, a professor at Ball State University, and research was conducted in the field by his students Melissa Castor and Andrew Castor. Jeffry Grigsby and Richard Fluegeman conducted specialized laboratory analysis on samples taken from a few of the boulders. Moreover, Edwin Crespo-Torres was hired by the city of Ponce to analyze the human remains recovered by the original excavations, which are deposited in the museum of the park. The results of both of these studies are presented in this volume, as well. After learning about the existence of the curated burials during his collaboration in our 2003 field season, William Pestle developed an interest in conducting bone chemistry analysis on them. His chapter in this volume presents the preliminary results he obtained from

The Archaeological Project of the Ceremonial Center of Tibes / 59 a pilot project directed to determining the feasibility of the project for use on the rest of the collections and the potential rate of success. It is important to stress here that while most of the fieldwork conducted to date by the project has been limited due to the detailed data being collected, the work of these specialists has provided invaluable data. This information, when combined with the meticulous field techniques used, has produced fine-grained results not normally found in the Caribbean (see Oliver 2000 for an exception).

4 Geophysical Prospection at the Ceremonial Site of Tibes, 1998Ð2001 Daniel Welch

This chapter discusses the methods and results of two geophysical surveys performed at Tibes over the 1998 and 2001 field seasons. The 1998 season represented the pilot project to test the effectiveness of one particular technique. Since the initial results were quite positive, the 2001 season greatly expanded on the 1998 season by incorporating two additional methods and the surveying of a much larger portion of the site.

Geophysical Prospection and Caribbean Archaeological Research Much has been written in support of the discipline embracing geophysical prospection in archaeological research. Kvamme (2001) points out the two major reasons most often cited: economics and conservation. As archaeologists know, field research is expensive. Once a site is located, traditional methods dictate a systematic program of shovel testing for location of archaeological deposits and then characterization of those deposits based on the recovered material. Depending on the nature and size of the site this initial exploratory phase may in itself require a field season. While it is certainly necessary to recover some material to assign identification to a deposit, shovel testing can be a poor way to locate deposits in the first place. A grid of shovel test pits will likely have quite a bit of unsampled terrain between them. It is often just chance that important deposits are located. On sites where conditions are amenable, geophysics provides the fastest, most complete way to locate deposits so that they can be tested (Clark 1990; Gaffney and Gater 2003;

Geophysical Prospection at the Ceremonial Site of Tibes / 61 Heimmer and De Vore 1995; Kvamme 2001; Witten 2006). This directly translates to a more effective use of excavation dollars. As for conservation, modern archaeologists typically approach a site with a set of clearly defined hypotheses to test by excavation. The conscientious researcher understands that excavation is destructive and wants to answer his or her research questions while minimizing damage to the site. The researcher must balance two major concerns when conducting excavations: first, minimizing damage to the site as a whole, and second, gathering enough data to answer the research objectives. Of course, the goal would be to excavate only those areas that contain the research material. Being nondestructive and noninvasive, geophysical prospection is thus a perfect tool to locate areas for excavation that have the highest potential to address the research questions at hand (Kvamme 2001). This is often the case with sites where the features are identifiable in the geophysical data. Good examples of this would include many historic structures, graves, and prehistoric sites with a known site structure. The concerns of economics and conservation are especially valid in the Caribbean. Many geophysical surveys have been conducted offshore with the aim of locating shipwrecks (Bequette 1991; Deagan 1983; Smith 1990). Given the size of a search area, uncertainty as to exact location, and difficulty of “test pitting” under water, remote sensing is vital to any marine project focusing on shipwreck recovery. In reality, geophysical prospection is the only viable way to locate these deposits. In each of the cited cases, towed magnetometers were used to locate the iron artifacts that would signify the ship’s location. After finding the wreck, standard underwater archaeological methods were employed. For terrestrial projects in the area, surface collection is still the preferred site discovery technique and geophysical prospection is used more for analyzing site structure (Curet et al. 2005; Fairbanks and Marrinan 1983; Gilmore 2008; Klingelhofer 2008; Roosevelt 1990). Much of the work has tended to focus on historic sites from early European settlement and on plantation sites in projects seeking to understand early European economic exploitation of the area (Fairbanks and Marrinan 1983; Gilmore 2008; Klingelhofer 2008). This is somewhat understandable given that historic sites with metal artifacts and sizable structures of nonperishable materials are more readily detectable with geophysics. Often the critique of geophysical survey for prehistoric sites is that the features are too ephemeral. While it may be true that some feature categories are easily masked by background conditions, the techniques are so varied and respond to so many different properties that the use of geophysics is appropriate more often than not (David et al. 2008). The chief goal of the geophysics at Tibes and of many geophysical surveys in general is to fill in the gaps left by traditional testing methods. At Tibes, for example, the site area was sampled during the 1995, 1996, and 1998 field seasons

62 / Daniel Welch with shovel tests dug to sterile soil. These tests were placed at the intersections of a 20-×-20-m grid and thus left a lot of unknown space between them. Since one cannot reasonably expect to shovel test the entire site, geophysics was planned to even out the picture by providing literally hundreds of thousands of additional data points within the shovel test grid. It was hoped that the data from the geophysics and the shovel testing as well as the full excavation would be complementary and that the geophysics would aid the overall project by mapping features. These features could then be singled out for excavation and since the excavation would be placed with less guesswork, more efficient use of time and resources would result, thus saving time, money, and aggravation. Furthermore, while excavation of diagnostic artifacts is certainly necessary to elucidate the cultural history of the site, geophysics sometimes allows the analyst to understand the geologic environment and the depositional history of a site. This primary information would be of great use to the parent project in reconstructing the development of the site.

The 1998 Pilot Project Electrical resistivity was chosen for the pilot project survey for several reasons. The primary reason was the nature of the features to be detected. Resistivity is excellent for detecting areas of compacted sediment or stone in moist sediments (Clark 1990; Heimmer and De Vore 1995; Weymouth 1986). The equipment is fast, easy to use and transport, and relatively durable when compared to other instruments. The presence of a sprinkler system with an unknown amount of ferrous metal and nearby high-tension power lines might have limited the effectiveness of magnetometry, and limited access to ground-penetrating radar equipment made GPR an unattractive candidate at the time. Tibes’s status as an archaeological park had a positive and negative impact on the survey. Low, maintained grass cover and a nearly flat ground surface made faster survey possible. There were serious concerns, however, that the increased soil moisture around sprinkler heads would produce pattern geometries in the form of low-resistance circles that could be mistaken for archaeological anomalies. In an attempt to avoid this issue, care was taken in the field to map the location of sprinkler heads. Furthermore, Tibes has been subjected to extensive previous excavation and reconstruction work. The effect that this previous work would have on survey results was another unknown factor. Electrical resistivity works by sending low levels of electrical current into the ground and measuring the resistance to that current caused by buried features. Typical systems introduce current with a series of metal probes. The resistances are read in units of ohms and are converted to units of resistivity (ohm-meters) after being corrected for probe spacing. Depth of investigation is selected by varying the distance between probes (Clark 1990). Tibes’s features were known from excavation to be fairly shallow, so a probe spacing of 50 cm was selected. Resistivity

Geophysical Prospection at the Ceremonial Site of Tibes / 63

Figure 4.1. Geophysical survey areas during 1998 pilot project (north at top).

is a volumetric measurement, so the reading would be affected by buried features within the top 50 cm of material. An electrical resistivity survey is conducted by dividing the survey area into a number of discrete grids. The instrument is then moved over the grid and readings are taken at evenly spaced intervals along evenly spaced survey lines; the higher the reading density, the finer the resolution. The original survey design called for a total of 19 20-×-20-m grids (Figure 4.1). The master site grid established in 1995 was used in order to avoid spatial error. The sample was not randomly selected but instead was concentrated in a rough rectangle around the central Plaza Principal and adjacent structures. A random sample would surely have included grid squares falling within plaza or ball court areas. Because the surfaces of these areas are extremely compact, they would have been difficult and time consuming to survey. Furthermore, it is likely that any remains of house floors would have been destroyed during ball court construction. As house sizes were expected to be no smaller than 3 m across, it was decided that a sample and traverse interval of 1 m would be an effective compromise between survey speed and resolution. Each survey grid would then comprise a to-

64 / Daniel Welch tal of 400 individual measurements. These measurements were recorded by hand in the field and entered into a computer for quick examination. Computer display while in the field was essential because it allowed the data integrity to be checked on site. Any error could then be corrected by resurveying. As the survey called for limited test excavations to evaluate the technique and identify anomalies, preliminary computer analysis was critical. Extreme high- and low-resistance anomalies noticed during actual survey were flagged for possible probing with a shovel test or controlled excavation. Any patterned geometries such as linear features, circles, or squares would be considered for excavation with a 2-m by 50-cm test trench. If the anomaly was found to be of archaeological importance, it was left unexcavated and reburied for careful investigation later. The cultural deposits at Tibes were known from previous excavations to occur at a shallow depth, typically 20 to 100 cm below surface. The instrument was configured to prospect at a depth of 50 cm. This depth was selected because it was deep enough to minimize modern disturbance while shallow enough to detect either a house floor directly or its effect on the immediately underlying sediment. A Gossen Geohm 3 soil resistance meter with probes configured in the Wenner array was used. The very common Wenner array includes four probes aligned in a straight line with the outer probes supplying the current and the inner measuring the potential. Probe spacing is equal to the prospecting depth. The steel probes were secured to a wooden frame so as to minimize any error caused by “leapfrogging” individual probes and changing their distance with respect to each other (Clark 1990). Due to weather and minor instrumentation difficulties, survey of 13 of the 19 planned survey grids was completed (Figure 4.1). Grids on the extreme eastern edge of the site were located in areas of heavy bush and were left undone. Grids 18 and 19 were partially surveyed due to vegetation and steeply sloping ground. After extensive processing in both Geoplot (Geoscan Research) and Surfer (Golden Software) software programs, examination of the final data was performed with an aim toward identifying possible cultural features. The final maps of each data set were interpolated to a measurement and traverse distance of 50 cm. This conservative approach was taken to ensure that any identified features were of possible cultural and not computer origin. In all map depictions, higher resistance appears as white and lower resistance as black. Each of the three data sets will be discussed separately below with reference to possible archaeological features. The findings from test trenches and shovel pits will also be discussed.

North Data Set The data set labeled “North” was the smallest contiguous group of grids sampled. It was composed of grids 1 and 2 and was partially in pastureland in its northeast corner. The pasture in this area was cut so that readings could be taken. It was hy-

Geophysical Prospection at the Ceremonial Site of Tibes / 65

Figure 4.2. Two depictions of the 1998 North data set of electrical resistivity; bottom view with indicated anomalies. Images are 40 × 20 m.

pothesized that the perennial burning of the pasturelands would release ions that would produce an overall depression in resistance values. Measurements recorded in areas of pasture were carefully mapped so as to test this theory. Due to its close proximity to the permanent datum point, we are confident that the corners of the North data set match those of the 1995 site grid. No shovel pits or test excavations were performed in this area. Each of the two grids was sampled after a day of heavy rain. Probe penetration was good, as there seemed to be a low density of near-surface rocks. Instrument readings were stable throughout and the array was not often deviated from the transect orientation of 90°. The most prominent anomaly of the North data set is the large circular area of lower resistance in the central area of the southern edge (Figure 4.2). It appears to be approximately 10 m in diameter and has a lower-resistance linear feature trailing off to the north. The shape of the anomaly strongly suggests cultural origin, but the fact that it is of lower resistance with respect to the surrounding area runs counter to the hypothesis that hard-packed domestic areas will exhibit high resistance. However, the low-resistance area may actually reflect the less-compacted edges of a dwelling or the trench dug to receive wall posts. The burning of the pasture produced no apparent effect. Immediately to the west of this horseshoe signature is an area of very high readings. After consultation with Lee Newsom, this is believed to be the signature of a mature algarrobo tree. This very large tree was responsible for the L-shaped area of no data nearby. The algarrobo is known for dense wood and has an especially dense and knotty root structure (Newsom, personal communication 1998). Such a root structure would cause the elevated readings.

66 / Daniel Welch An excellent candidate for a house feature is the square anomaly in the northcentral area. This feature appears to be 5 to 7 m across with a branch of two possible footpaths leading northeast and southeast emerging from its eastern side. To the southeast of this feature is a more diffuse area of higher resistance, which could signify an outdoor activity area. The dip in resistance in the center of the feature is likely the result of the Wenner array. It is common to read a drop between two peaks (Clark 1990). This results in a “donut”-shaped anomaly and should not be taken to mean that the center of this feature is actually of lower resistance. It is noted, however, that without excavation we cannot be sure.

East Data Set The six grids of the East data set represent the largest continuous area surveyed during the 1998 season. These grids ran along the entire eastern edge of the central Plaza Principal (Structure 6) and the Plaza de Estrella (Structure 7). Two grids were also planned to the south of the Plaza Principal with the theory that the area between it and the Batey de Herradura (Structure 2) would yield evidence for high-status houses based on close proximity to ceremonial areas. These grids were unremarkable during survey and only one shovel pit was planned to probe anomalies. By the time this survey was conducted, late in the field season, other excavations had conclusively proven that high-resistance readings were concentrations of stones and not packed-earth floors. An area of stable high resistance was noted in the northeast corner of grid 11 and can be seen on Figure 4.3a–b. From a 1996 test excavation in the area, we knew this deposit to be cobble overburden from field clearing and ball court construction or an indication of a Portugués River paleochannel. Resistivity allowed the deposit to be mapped to a large extent. Unfortunately, irreconcilable contrasts between the edges of this grid and its northern neighbor obscure its northern limit. The computer-generated anomaly seen in the low-pass filter display may more closely approximate this feature’s actual size (Figure 4.3c). A low-pass filter removes the effect of high-frequency changes caused by archaeology and is useful for examining evidence for geologic interference. Other areas of elevated values are located in the southeast area of grid 16 and in grid 14. High values in grid 14 may be the result of overburden or close proximity to the ball court’s “sidewalk” that caused better drainage. The most interesting aspect of possible anomalies in the East data set is their orientation. Features appear as roughly rectangular with a long axis running northwest to southeast. Although it is difficult to interpret on the paper maps, linear features with the same orientation are observed in the southeastern section of the data set. Upon close examination, these anomalies appear to have an area of concentrated higher values to the southeast and a more diffuse area to the northwest. Although the survey resolution is too coarse for this interpretation to be absolutely certain, the anomalies suggest a roughly rectangular habitation with

Figure 4.3. Three depictions of the 1998 East data set of electrical resistivity: (a) electrical resistivity with indicated anomalies; (b) electrical resistivity; (c) low-pass filter display.

68 / Daniel Welch a clearly demarcated house in the southeast and a more spread-out activity area in the northwest. During the survey of grid 15 an area of moderately high resistance was noted in the north-central section of the grid. This anomaly tapered off very quickly and was chosen as a possible candidate for a house floor. A 2-×-.5-m test trench was centered over it, but due to time constraints this was discarded in favor of shovel testing. The shovel pit (indicated on Figure 4.3a in grid 15 with a small circle) produced a moderate concentration of fist-sized cobbles at 12 cm and a low density of ceramics at approximately 15 cm. Possible ground-stone lithics were also collected. The excavation was halted at 48.5 cm because large concentrations of rounded cobbles made further work impossible. As can be seen from the final map, the location of the shovel pit is at the northwestern edge of a possible cultural anomaly. The recovered material may then represent floor sweepings from a house. The close proximity to a tree may also have been responsible for the elevated readings in the area immediately around it. The conclusion from this shovel pit is that there is little clear evidence for house floors at this location.

West Data Set The West data set is made up of five grids running along the western edge of the central portion of the site to the western bush line (Figure 4.4). Grids 18 and 19 were surveyed as close to the bush line as possible, though it should be noted that the western edge of the cleared portion of the site runs roughly southwest with respect to the survey transect lines. Test excavations from the 1995 and 1996 field seasons indicated the presence of domestic deposits in this area and the main project excavation of the 1998 season was conducted in the southeast section of grid 18. These grids (with the exception of grid 4) were the first to be surveyed so as not to impede excavations. The West data set received the most intensive field computer processing. An effort was made to hand-concatenate grids, as some anomalies could be seen crossing grid edges. This data set also received the most intensive test excavation. Three trenches and a shovel pit were dug to test different classes of anomalies. A shovel pit was dug over a small area of extreme high resistance in grid 7. It is indicated on Figure 4.4a with a small, white enclosure. The measured value of approximately 250 ohms was nearly 10 times higher than the normal, most frequently observed measurements. There was concern that this spike may have been caused by instrument error, but computer display showed that the small region in the “crook” of the Plaza Principal’s western sidewalk exhibited much higher values throughout. An extremely high concentration of cobbles was found. Many of these stones were of similar size to those that make up the nearby sidewalk. A small amount of ceramic, coral, and historic material was recovered. According to

Geophysical Prospection at the Ceremonial Site of Tibes / 69

Figure 4.4. Two depictions of the 1998 West data set for electrical resistivity: (a) electrical resistivity with indicated anomalies; (b) electrical resistivity.

the head groundskeeper, the ball court and plaza sidewalks were often underlain by a stone foundation. This may indicate that the current reconstruction and consolidation done during the original excavation is incorrect in this case (González Colón 1984). The shovel pit was halted at a depth of 85 cm. The sediment at this depth was well-sorted beige sand, which has proven at Tibes to be sterile. Unit OP19A, the first test trench, was oriented east–west over an area of high resistance in the northwest corner of grid 19. The location was flagged during survey as a possible test excavation site. Unfortunately, the unit proved to be located on nearly the exact location of one of the 2-×-2-m exposures from the original work by the Sociedad Guaynía. A large amount of historic artifacts and mixed stratigraphy confirmed the disturbance. The most notable aspect of Unit OP19A in terms of the resistance values was not the fact that it was disturbed but that it contained a high density of cobbles and some large boulders. These boulders first appeared at less than 20 cm below ground surface and continued beyond the maximum depth of the unit, which reached just over 60 cm. The size of the rocks and the fact that many smaller cobbles and pebbles were found among them accounted for the high

70 / Daniel Welch spike. This evidence, combined with the shovel pit results, shifted the hypothesis slightly in that the areas of highest resistance were no longer considered likely house-floor locations. Instead, further test excavations were planned for areas of moderate resistance. Field computer display of the western survey area showed several areas of “diffuse moderate resistance” that seemed to be connected by a network of relatively straight linear features. One of these areas was selected for Unit OP19B. The unit was placed so that it would be half over the feature and half over what appeared to be vacant terrain. It proved extremely difficult to correctly place small test trenches in accordance with the survey grid, and deviation of approximately 50 cm to 1 m from the display maps can be assumed. Unit OP19B was a 2-m by 50-cm test trench oriented north–south. It was dug by natural stratigraphic layers and was to be halted at a sufficient depth to confirm the resistance reading. At a depth of 30 cm below ground surface, a possible hearth was found protruding from the western wall on the southern half of the unit. It was a dense, circular concentration of stones 70 cm wide and contained numerous charcoal flecks and fragments of shell. The area around the hearth and the rest of OP19B was almost clear of stones, though charcoal flecks were evident. The unit was subdivided into locus A (north half ) and locus B (south half ). Excavation continued in locus A while the possible hearth and locus B were left unexcavated. Concentrations of charcoal were found, as well as scatters of bone, shell, and ceramics. Unit OP19B clearly showed evidence of an activity area. Radiocarbon analysis of charcoal from units expanded from this feature during the 1999 season dated it at a.d. 1015–1205 (cal 2s), placing this deposit near the end of the Elenan Ostionoid period (Curet et al. 2006; Newsom and Curet 2000; see also Curet, this volume). Unit OP19C was located 2 m to the southwest of OP19B and was planned so as to straddle the trail that appeared to connect to the domestic area discovered in OP19B. The unit was a 2-m by 50-cm test trench oriented east–west and was dug by natural levels to a depth sufficient to confirm the resistance measurement. Shells and ceramics were found beginning at level 2 (25 cm below datum [cmbd]), as well as charcoal. A greater overall concentration of charcoal was evident in the western half of the unit. After 25 cmbd, the sediment became noticeably harder. After the excavation of level 3, it became clear that a trail was present. The sediment in the central portion of the unit was much harder than on either side and was nearly devoid of artifacts. In contrast, the ends of the unit were littered with small ceramics, shell, and charcoal. The trail sediment was a light–medium brown, while on either side the sediment was dark, greasy, organic brown-black. Water was sprayed over the level, and the differential drying characteristics of the areas confirmed the trail location. As expected, the trail ran roughly north–south and was approximately 1 m wide. The unit was subdivided into locus A (western half ) and locus B (eastern half ). Excavation continued in locus A in order to define the trail.

Geophysical Prospection at the Ceremonial Site of Tibes / 71 The hollowed-out portion of locus A continued through levels 4, 5, and 6, and eventually reached a depth of 70 cmbd. The deposit was characteristic of a midden fill and had ceramics, shell, bone, charcoal, and dumped ash.

The 2001 Project Since the 1998 season showed that geophysics had some potential at Tibes, it was decided that the 2001 season would incorporate a much more intensive survey aiming to cover the majority of the site with electrical resistivity, magnetic gradiometry, and ground-penetrating radar (GPR). These three tools were selected for complementarity reasons. It was hoped that full-coverage survey with resistivity and magnetic gradiometry would be able to detect domestic areas with greater certainty. Resistivity would be sensitive to house floors and middens, while the magnetic gradiometry should pick up associated hearths and burned features. GPR was also selected because it would be able to differentiate features at differing depths. The resistivity unit selected for the 2001 season was the Geoscan RM15. This unit is built specifically for archaeology and features automatic data logging. This allowed a much greater rate of collection than was obtained in the 1998 season. As a result, resistivity data were taken every 50 cm along transects spaced 50 cm apart. This represents a dramatic increase in data density over the 1998 pilot survey. Each 20-×-20-m grid now had 1,600 data points as opposed to 400 in 1998. The other difference in the resistivity methods in 2001 was that the equipment was configured in the twin-probe array. The twin-probe array is composed of two roving probes and two stationary probes. The stationary probes are kept 30 times the roving probe spacing away from the roving probes. This method is superior to the Wenner array used in 1998 because it is less likely to introduce artifacts into the data (Kvamme, personal communication 1999). A total of 51 20-×-20-m grids were surveyed with resistivity. Magnetic gradiometry was chosen as a strong complement to resistivity. Magnetometry is a passive technique that measures the variations in the strength and flux of the earth’s magnetic field (Breiner 1973). These variations are caused by buried objects or by alterations in the magnetic properties of the soil. Intensive burning, for example, will significantly increase the magnetism of an area (Heimmer and De Vore 1995). Burned areas were thus the main targets of the magnetic survey. It was hoped that full-coverage resistivity would identify house floors while the magnetic gradiometry would find the hearths. Magnetic gradiometry is a specific refinement of magnetometry that is used for archaeological prospection and other applications that require detection of ephemeral targets. Magnetic gradiometry involves the use of two magnetometers. One magnetometer is used to record the data across the site. The other records the

72 / Daniel Welch natural variation in the earth’s magnetic field throughout the day. This natural variation is called the diurnal effect and can significantly skew survey data if it is not taken into account (Clark 1990). One added complication was the presence of igneous rock on the site. This creates a great deal of background “noise” and is a factor that must be taken into account on volcanic islands (Gilmore 2008). The instrument used for the survey was the Geometrics G-858 Cesium Vapor gradiometer, which incorporates two vertically spaced magnetometers. The lower one records the variations caused by buried features and the total magnetic field. The upper one records just the total magnetic field. This allows the data from one magnetometer to be subtracted from the other, thus leaving signals only influenced by buried features. The magnetic gradiometer data were spatially corrected in Geometrics MagMap2000 software and then outputted to Golden Software’s Surfer for interpretation. The magnetic field strengths were mapped to a color table. Display colors were varied in order to accentuate weaker signals and to winnow out the very strong spikes in the data. These spikes were likely the result of metallic debris or sprinkler-head fittings. Data were taken every 10 cm along transects spaced 50 cm apart. Forty-seven 20-×-20-m grids were surveyed with magnetics. The third method chosen was GPR. Like electrical resistivity, GPR is an active technique in that it induces energy into a material and records that material’s response. GPR sends rapid pulses of low-level, high-frequency electromagnetic energy into the ground and measures the time and strength of any reflection. Depth of investigation is set by varying the reflection listening time, or time range. Data are recorded along evenly spaced profiles at constant intervals. Horizontal planview slices are interpolated at varying time depths in order to generate a map of reflection amplitudes. Analysis of the GPR data from Tibes is ongoing and will be completed at a future date. The 2001 data were divided into two areas for interpretation. The North data set extends north from N200 (Figure 4.5). The South data set encompasses the surveyed area south from N200.

North Data Set On sites with large, easily recognizable architecture, the geophysical analysis has only to key into pattern geometries. However, when there is no large architecture present or on many prehistoric sites, the analysis must look for concordance between data sets to determine likely areas for excavation. The working theory was that elevated resistance paired with a nearby magnetic high would indicate a living surface and associated cooking area. Figure 4.5a–b shows magnetics (a) and resistivity (b) for the northern portion of the site. The north edge of the Plaza Principal is visible at the bottom of the images. Strong anomalies are displayed in white for the resistivity and as a

Figure 4.5. Three depictions of the 2001 North data set: (a) magnetic gradiometry; (b) electrical resistivity; (c) electrical resistivity with result of map subtraction (see text).

74 / Daniel Welch paired white/black for the magnetics. Anomalies that appeared in both data sets at the same approximate location were chosen as “high confidence.” High-confidence anomalies are shown as white circles. Test excavations planned over some of these anomalies turned up high concentrations of cobbles. Many of the stones and cobbles that make up the ball courts are highly magnetic and thus produce strong signatures on the magnetometry results. The jumbled appearance of the southeastern portion of the star-shaped plaza (Plaza de Estrella) in the magnetic data as well as the diffuse high-resistance readings for the same area may indicate that the present reconstruction of that plaza is incomplete and it is missing a star point. Figure 4.6a shows some detail of a magnetic anomaly from the southern edge of Batey 9. The outlines of the ball court are visible as a result of the magnetic stones in the rows. A test unit was dug to test the pair of strong anomalies enclosed by the white circle. While the excavation did not yield any evidence of domestic activity, two large boulders were found very close to the surface. At first glance one of them appeared to have been extensively worked, but later analysis seems to indicate that the peculiar facets on it are natural in origin. The stone was volcanic in nature and was thus likely the source of the anomaly. Figure 4.6c shows this object in situ. Interestingly, GPR data collected in the area also showed some circle-shaped anomalies in the center of the ball court a little to the north of the magnetic anomaly shown in Figure 4.6c (Figure 4.6b). Since GPR data processing and interpretation are still ongoing, these anomalies have yet to be tested with excavation.

South Data Set Figure 4.7 shows the magnetic and resistivity data from the southern section of the site. The solid white circles show areas of concordance between the two data types. A close examination of those indicated areas in the magnetic data will show that each magnetic anomaly is a paired positive/negative peak. A single, paired positive/negative is likely the result of a single object. Excavations of like magnetic anomalies from the North data set indicated clearly that shallowly buried boulders showed as a single strong anomaly. Those locations also show a very strong spike in the resistivity data. Again, based on excavation experience from the northern areas, this is consistent with buried boulders. One interesting feature in the magnetic data is the anomaly at 115, 90 (Figure 4.7a). It looks to be a single, paired positive/negative peak of the type that would indicate a single boulder. However, there is no corresponding signal on the resistivity data because it falls right on the edge of the surveyed area (Figure 4.7b). We may be seeing some indication of this in the slightly elevated resistivity readings along that edge. This anomaly falls inside of a ball court. Excavations of a similar anomaly in Batey 9 revealed the large boulder shown in Figure 4.6c. While lithic analysis seems to indicate that the faceting on the boulder is natural, perhaps its placement is not and could be indicative of a pattern.

Geophysical Prospection at the Ceremonial Site of Tibes / 75

Figure 4.6. Anomaly in Structure 9: (a) magnetic anomaly; (b) round GPR anomaly; (c) source of anomaly.

Discussion and Reevaluation Further test excavations showed that the selection criteria used for identifying anomalies for testing were flawed. The magnetic data proved to be unreliable due to the presence of so many magnetic cobbles. A reexamination of the resistivity data should be done with an aim toward identifying weaker resistance areas. Since

Figure 4.7. Two depictions of the 2001 South data set: (a) magnetic gradiometry; (b) electrical resistivity.

Geophysical Prospection at the Ceremonial Site of Tibes / 77 excavation demonstrated that the strong anomalies were concentrations of rock, weaker-resistance anomalies may be the signatures of house floors. Furthermore, areas of medium to high resistance were also found to be large, dense concentrations of rocks. In order to better understand the distribution of areas of high stone concentration, and try to separate those signals from the rest of the data, specialized processing was done on the resistivity data. All of the resistivity data were low-pass filtered. A low-pass filter is a technique that shares many commonalities with image processing. The idea is to move a large window over the data and to run a mathematical algorithm to smooth out spikes (Russ 1995). The low-pass map would be used to locate the rock concentrations and thus identify areas that may be problematic for both resistivity and magnetics. The presence of a great amount of near-surface rock is certainly skewing the resistance data and making it extremely difficult to separate the archaeology from the background. Furthermore, it is clear that a majority of the rock types at Tibes are volcanic in origin and are thus magnetic. The detection sensitivity of a magnetometer falls off rapidly with respect to object distance from the sensor. As a result, areas with shallow (less than 50 cm below surface) rock will pose a significant challenge to the analysis of the magnetic data. Resistance data that is low-pass filtered will enable two types of analyses. First, a pattern to the distribution of rock may shed light on the depositional history of Tibes sediments. Any archaeological features should be spatially small and smoothed out by this filter. Low-pass filtering is a fairly common step in resistivity data processing because it shows how much of the signal may be the result of gradual changes in underlying geology. The result of a low-pass filter is a map that shows only the broad changes in the data. This can allow the analyst to better understand the degree to which anomalies may be geological. The low-pass filtered data are shown in Figure 4.8. High-resistivity areas are shown in white, low in black. Although the low-pass filter was performed in order to remove the influence of the gradual signals from the geology, it did highlight some of the interesting geomorphic processes that have been at work at Tibes. The high values form bands that trend from the northwest to the southeast across the site. It is likely that the low-pass data are imaging the Portugués River’s paleochannels. Furthermore, the cobbles that were found in many of the test excavations appeared to be rounded river cobbles. It may be that Tibes was built on a terrace formed by a series of gravel bars. A geomorphological study of the area would test this theory. The second reason for the low-pass filter map is to do map subtraction. In order to highlight the small changes that may be archaeological features, the result of the low-pass filter is subtracted from the original data set. That removes the gradual changes and more clearly shows the archaeology. The results of the map subtraction for the 2001 North data set are shown in Figure 4.5c. Immediately noticeable in the Figure 4.5 data is the number of small anomalies scattered east of the

78 / Daniel Welch

Figure 4.8. Depiction of the 2001 low-pass filtered data for the whole site.

Plaza de Estrella. These small signals are likely the result of rock concentrations and, while there may be archaeological features in this area and in the data, the archaeology is masked by the signals. Concentrations here are so great and so varied that even the subtraction of low-pass filtered data was unable to completely remove their influence. What is required now is to choose anomalies for testing that fit the revised se-

Geophysical Prospection at the Ceremonial Site of Tibes / 79 lection criteria. These anomalies should have low–mid resistivity signals and not occur in an area of small, high-resistivity anomalies that may be rocks. These new anomalies will be tested with excavation in coming seasons and will, we hope, be able to productively guide future work at Tibes.

Conclusion Geophysics at Tibes remains an ongoing study. The data gathered over the 1998 and 2001 seasons have the potential to help the overall project accomplish its goals more quickly by identifying areas with higher potential for domestic features. The work is complicated by several factors. First are the features themselves. The house floors and pits are ephemeral, low-contrast features. Furthermore, they are not expected to present any clear diagnostic structure that would be discernable in the geophysical data. Second, the underlying geology and the magnetic nature of the stones add noise to the data and may obscure whole sections of the site. This is the case with the eastern edge of the site. Finally, there may be quite a bit of disturbance at Tibes associated with the original work and subsequent conversion to an archaeological park. In spite of these obstacles, geophysics has shown to produce meaningful results. Test excavations and a review of the data in light of what they find will allow the interpretations to evolve. These conditions highlight a key point in archaeological geophysics; that is, the need for ground truth calibration of data. Even a brief study of geophysical survey reports will yield scores of examples in which the data look like an x-ray of the ground. Roman fortresses and villas leap from the page in staggering detail. On many prehistoric sites, the geophysicist does not have the luxury of anomalies caused by big stone walls. Often the only detectable difference is a slight change in soil chemistry. What makes one blob a feature and another blob noise? That is the question that the analyst must confront. The surest way to produce a data set that is effective is to continuously reinterpret with excavations. The geophysical surveyor cannot just breeze in, collect data, and toss a map over his or her shoulder on the way back to the airport. There needs to be constant feedback between excavation and data results. If that is the case, then the analysis will evolve and change and only become more useful and valuable with time. The work at Tibes thus does not represent a static map. There is plenty of archaeology in the geophysical data. There is also plenty of geology. Feedback from excavation will teach us how to calibrate this particular data set and fine-tune it so that it becomes a more effective tool to understand the site.

5 Paleoethnobotanical Research at Tibes Lee A. Newsom

Paleoethnobotanical research at Tibes, a relatively early Caribbean civic-ceremonial center in southwestern Puerto Rico, was initiated in 1995 and is ongoing. The central focus of this research is to clarify the indigenous ethnobotany and contribute to a better understanding of the evolution of social organization at the site. The intrinsic dynamics of coupled human-natural systems, with shifting baselines and flux in resource abundance and availability, both natural and culturally mediated, are another focal point of this research. Tibes represents one of the earliest communities in the region to demonstrate a level of social complexity approximating a chiefdom. The site contains early and later components, specifically late Saladoid to early Ostionoid (ca. a.d. 600–1000) deposits, that encompass the period of transition from a more egalitarian form of social organization to one with some degree of social differentiation. Thus successive deposits at Tibes potentially contain evidence of key details associated with the emergence of a local Caribbean chiefdom, affording an opportunity to examine this development, including such examination from the perspective of the subsistence enterprise. This would encompass ethnobotanical aspects of subsistence and how they parlayed into wider trends in household and political economy, the latter perhaps including the control of subsistence goods (Hirth 1996; Johnson and Earle 1987). The paleoethnobotany thus may serve to illuminate these dynamics through the specification and understanding of the particular types, roles, and presence or availability of plant resources at the different levels of analysis and complexity ranging from individual households to the community at large, i.e., the corporate arena, and at particular points in time. One example would be evidence of plants involved with the appearance of feasting, as an activity potentially

Paleoethnobotanical Research at Tibes / 81 tied to emerging elites (Hayden 1995) and indicative of plant use transcending common household consumption. More broadly, could the development of occupational specialization in the region, including perhaps at Tibes, have involved biotic resources (e.g., someone in the role of a specialist using medicinal herbs) and could their control have been a potential stimulus to chiefdom development? Did emerging elite assume roles as entrepreneurs, undertaking responsibility for management of environmental risk through economic (in this case, food collection and storage) or other methods? The realities of biotic resource concentration, especially in view of social or geographic circumscription, as a coercive force in the aggregation of polities also remain to be explored. Considering these ideas and others relevant to the Tibes paleoethnobotanical research, I began by simply asking what plants were present and whether there was evidence to suggest cultivation, i.e., management. What were the basic elements of household economies, especially in view of the assumed trends in sociopolitical status over time? What were the relative contributions of wild vs. domesticated plants; of native vs. introduced taxa? What role did maize (Zea mays) play, if any at all, given its low profile archaeologically in the region (Newsom 2006)? Did any plants serve as staples, and if so, was staple-crop production focused or diffuse? To what extent were subsistence practices of all types sustainable and how might this bear on social organization? Is there any evidence, in fact, for feasting or control and manipulation of crops or other significant plant resources by elites? If so, did crop production parlay into some form of staple finance (Hirth 1996; Johnson and Earle 1987; Rountree and Turner 1998)? Could living in a relatively forgiving, i.e., productive, resource-rich subtropical environment allow for sufficient population sizes to encourage the growth of chiefdoms without a primary emphasis on agriculture, i.e., staple-crop production? Is there any evidence for food or other resource intensification practices, especially if they may have served as a means to support or justify an emerging elite? And just as guinea pigs (deFrance et al., this volume) were evidently bred and maintained in captivity, were special-use plants (e.g., narcotics, condiments) specially cultivated, and if so, for what reasons (e.g., foods reserved for feasting, foods reflective of social relations, plants used in ritual)? Was access to particular plant resources restricted to social segments within the community, i.e., exclusionary? Considering these questions, among others, the paleoethnobotanical research was initiated with select samples recovered during a series of systematic excavations briefly overviewed below. This effort emphasized samples intended for analysis of plant macroremains and pollen from a variety of discrete features, activity floors, middens, and bracketing deposits. At this juncture my interpretations are focused at the household or, perhaps more correctly, suprahousehold level (Curet et al. 2006), a necessary initial step in data collection and pattern recognition as a baseline toward addressing change and answering some of the broader

82 / Lee A. Newsom questions and details of social dynamics posed above. Indeed, distinctive plants and patterns of use at Tibes, as shall be seen below, suggest considerable potential to clarify trends in social evolution at the site. These data also illuminate local habitats and forest communities, which helps to conceptualize the ancient landscape and allows initial consideration of the types of biotic resources potentially available to the site’s inhabitants. This chapter provides a synopsis of the research thus far. Because of space limitations the presentation and discussion of the results are scaled to a minimum. For any who may have an interest, the original full-length chapter (Newsom 2009) is available as a pdf file directly from the author.

Paleoethnobotanical Background By now Caribbean archaeologists have generated a fairly secure timeline and anthropological understanding of the trajectory of human occupation and social organization for the region (Keegan 1996, 1997b, 2000; Keegan et al. 1998; Rouse 1986, 1992; Siegel 2004, 2005; Wilson 1990, 2007), along with an increasingly refined idea of subsistence–economic dynamics. Mobile or relatively sedentary hunter-forager-fisher-gardeners of the Archaic age were more or less succeeded by village-dwelling and seemingly relatively egalitarian tribal or segmentary societies early in the Ceramic age who probably had some form of informal leadership and who evidently relied extensively on root-crop horticulture. Socially ranked groups emerged later in the Ceramic age, at least in some areas, and presumably with true hereditary leaders. These societies in particular are ones in which plant production may have been more intensive and assumed a wider cultural significance. Ultimately, in the years and decades following European contact, several early chroniclers recorded details of Taíno subsistence and economic (or political economic) behaviors (Fernández de Oviedo 1851–1855 [1535]; Jane 1988; Lovén 1935; Pané 2001 [1505]); in general, household subsistence economy seems to have revolved around aquatic faunal resources and root-crop production. Some documents provide insights into agricultural activities at the time of contact, including the creation of formal space, mounded growing platforms, for crops (Fernández de Oviedo 1959a [1526]:13–14; Las Casas 1971 [1527–1565]:110; Sauer 1966:51–54). Others suggest the presence of home gardens and “special beds” (Sauer 1966:57) maintained in and around settlements (Dunn and Kelley 1989:117; Layfield 1995:150– 153; Sauer 1966:56–58). Any number of introduced trees and herbs that were sources of fresh fruit, condiments, medicines, dyes, fiber, inhalants, or narcotics were probably maintained in these settings (Newsom 2008; Nieves-Rivera et al. 1995; Rouse 1992:14). There are also suggestions that the Taíno may have managed particular protein sources, e.g., native rodents and turtles (Newsom and Wing 2004). Guinea pigs (Cavia porcellus) were among the introduced captive fauna now

Paleoethnobotanical Research at Tibes / 83 confirmed archaeologically by the late Ceramic age (Wing 2009). DeFrance et al. (this volume; deFrance 2009) have documented their presence at Tibes and hypothesize that they likely served nonculinary purposes such as offerings or in traditional curing and divination rituals. This idea may have some bearing on plant use at Tibes, something to which I return later. My central point in relating the above is that this reflection of increasing human social complexity through time and space necessarily implies different uses, perspectives, roles, and degrees of reliance on plant resources over the course of human occupation in the region, and my basic premise or hope for the paleoethnobotany, including that of Tibes, is that it can reflect back and inform us on the various inherent social dynamics, including changing economic behaviors. For example, in situations in which crop production entailed agricultural landscape engineering and capital improvements, and beyond that related to social feasting and obligations such as tribute goods that were comprised of surplus foodstuffs, we may infer some level of direction and control under the aegis of individual powerful leaders (Johnson and Earle 1987) such as sooner or later must have been associated with centers such as Tibes. To these ends, plant remains recovered and now identified from sites found throughout the region, but particularly Puerto Rico, provide an interesting and lengthy record of human ethnobotanical behaviors in the islands. It has become clear that plant cultivation, not simply “hunting and gathering,” dates to the outset of human occupation in the archipelago, and eventually some taxa factored into social realities of status and ritual, e.g., cojobilla (cohoba) and perhaps some others (Newsom 2008). These are all topics under consideration with regard to the paleoethnobotanical research at Tibes.

Southwestern Puerto Rico The earliest evidence for human occupation in southern Puerto Rico, including the general area surrounding Tibes (but not at Tibes itself ), is attributed to groups of the Archaic age, a period encompassing the first presence of humans in the region in general (see Rodríguez 1997). At the time we initiated our work at Tibes, very little previous paleoethnobotany had been conducted on Puerto Rico. In the southern region of the island where the site is located, El Bronce, nearby and roughly contemporaneous with the later Ostionoid component at Tibes, had been analyzed by Pearsall (1985). Her work provided some insights into fuelwood reliance for the period, with buttonwood mangrove (Conocarpus erectus, Combretaceae), the genus Coccoloba (Polygonaceae; coastal C. uvifera, sea grape, or one of several interior forest species [Liogier and Martorell 2000:49–50]), and croton (Croton sp., Euphorbiaceae) prominent among the taxa identified. These may have been preferred fuels, if not simply trees and shrubs that were locally abundant, thus readily available for fuelwood collection. Elsewhere, in the Cerrillos River drainage, analysis of small sets of samples

84 / Lee A. Newsom from three other Ceramic age sites—PO-21 (late prehistoric, also a historic component), PO-38 (El Parking site; Saladoid [Cuevas] and Ostionoid components, ca. a.d. 500–1200), and PO-39 (La Iglesia de Maraguez; Saladoid-Ostionoid, ca. a.d. 100–1000 [deFrance and Newsom 2005; Newsom 1987, 1990, 1992])—provided some limited indications of fuelwood use and reliance on edible plant resources. Most notable are two identifications from PO-38: papaya or lechosa (Carica papaya, Caricaceae) and Oenothera sp. (Onagraceae), evening primrose. The PO-38 papaya is the only archaeological record thus far in the Caribbean for the presence of this multipurpose plant (fruit, medicine), which originates in either Central or South America (Newsom and Wing 2004:154). Seeds from evening primrose, a mildly narcotic herb, had also been (actually was originally) identified from El Fresal, another Ceramic age settlement on Puerto Rico to the northeast of the Ponce area (Barrio Cuyon, Aibonito [Newsom 1988]). This taxon presents a puzzle because it no longer exists as part of the island’s flora (Liogier and Martorell 2000), but it nevertheless has now been identified from a total of five Ceramic age sites on Puerto Rico (deFrance and Newsom 2005; Newsom and Wing 2004), including in the southern region at PO-38 and El Fresal, as well as Tibes. More about this below. Finally, at Archaic age Maruca (Newsom and Wing 2004; Pagán-Jiménez et al. 2005) reliance on tropical fruits and edible rootstocks is suggested, perhaps including some under cultivation. Maruca may therefore be part of the growing body of evidence for the early cultivation of starchy staples. Archaeobotanical data from Puerto Rico, including the southern part of the island, comprise a significant portion of the taxa now recorded for the Caribbean as a whole (Newsom 2008). A number of tropical fruits were utilized by the original inhabitants as comestibles or for other purposes, as were maize, certain tropical root crops, and several herbs, beginning at least as early as the Archaic age. Maize here as elsewhere in the region seems never to have achieved the status of a staple crop prehistorically (Newsom 2006, 2008). Instead, the tropical root crops may well have served that role, based at least on the ubiquitous presence in Ceramic age sites of material culture commonly associated with manioc processing, now corroborated to some extent by microbotanical evidence such as from Maruca and by isotopic analyses.1 We now also have a reasonable idea of fuelwood preferences and have an indication of the spatial variability of fuelwood use across the island in prehistory. Aside from what I noted earlier regarding El Bronce, I will simply refer the reader to the discussion of carbonized wood data provided in deFrance and Newsom (2005). In general, differences in wood use across the island appear to have related mainly to differences in forest species and diversity (northern moist vs. southern dry limestone regions, and the central mountains). It seems, however, that the Tibes wood assemblage, especially in comparison with El Bronce, reflects also something of

Paleoethnobotanical Research at Tibes / 85 the specific nature of the site as a developing civic-ceremonial center, a point to which I return later.

Natural Environment in Brief Tibes is located in the subtropical dry forest region of southern Puerto Rico about 8 km inland from the south coast along the banks of the Portugués River (see Figures 1.1 and 1.2), near the present-day city of Ponce. The area experiences significantly lower annual precipitation and regular moisture deficits compared to the windward northern and northeastern portions of the island (Murphy et al. 1995). The average annual rainfall in the southern dry zone is around 600 to 1,000 mm, depending on the specific location and stochastic variation from one year to the next (Lugo 2005:402; Murphy et al. 1995:179–180). Bioclimatic conditions in the Ponce area equate with the “subtropical dry forest life zone” of the Holdridge system (Ewel and Whitmore 1973; Murphy and Lugo 1986). The predominant vegetation is classified as dry to very dry subtropical forest, and it reflects an evolutionary capacity to survive under perpetually arid conditions and periodic severe moisture deficits in its overall physiognomy and phenological patterns (Ewel and Whitmore 1973; Murphy et al. 1995). Guánica Forest, just west of Ponce, is today one of the best remaining examples anywhere of subtropical (and tropical) dry forest ecosystems, and it is considered to be representative of the original vegetation of the arid southern coast and lowlands of Puerto Rico (Murphy et al. 1995). Six distinctive plant associations variously make up the basic fabric of this forest ecosystem mosaic, spanning the shoreline to more upland settings and depending on local ambient soil and moisture conditions (Murphy et al. 1995). Three coastal plant communities include mangrove forest, salt flat, and “beach thicket.” Farther inland the three other major associations range variously into the higher terrain such as in the immediate area of Tibes, including scrub forest on sandy, shallow soils and limestone outcroppings; semievergreen to evergreen forest on sites with deeper alluvial and colluvial soils; and deciduous forest under more or less intermediate conditions (Lugo 2005:416; Murphy et al. 1995:186–187). Together, slope, sun aspect, soil depth, and moisture availability are the primary determining factors in the spatial extent and ultimate physiognomic expression of these upland vegetation forms on a continuum spanning deciduous through evergreen (Lugo 2005:416; Murphy et al. 1995:186–187). In addition, interannual variability in rainfall exerts a strong influence on the physiognomy: “when annual rainfall is greatly reduced, the forest acquires a more deciduous aspect” (Murphy et al. 1995:192). This has clear implications for understanding the forest history of the region. The deciduous forest formation comprises the greatest areal extent of

86 / Lee A. Newsom forest cover today, dominating most upland areas in the southern region, including Guánica. The dry forest vegetation in Puerto Rico tends to be low in stature and relatively low in species composition compared to moist and wet forest associations on the island (169 species of trees occur in Guánica’s deciduous forest [Murphy et al. 1995:188]). Thorny and microphyllous (small leaves) species are common (Lugo 2005). The timing of vegetative growth and the activity of the vascular cambium (i.e., wood and bark formation) are strongly correlated with the seasonal cycles of moisture availability, with peaks during the two wetter periods; likewise, flowering is positively correlated with the wetter months of the year (Murphy et al. 1995:195). In contrast, overall fruit production of tree species in Guánica has been found to be relatively constant and abundant throughout the year with little seasonal pattern (Murphy et al. 1995:195). Indicator trees associated with the subtropical dry forest life zone on Puerto Rico (Ewel and Whitmore 1973:16–20) are listed in Table 5.1, along with a list of key arboreal taxa (Importance Values [IV] >10.0) associated specifically with the deciduous forest association based on tree surveys and census data from Guánica (Murphy et al. 1995:Table 7). These lists highlight some of the principal components of the local forests, and it is not surprising therefore that several of the taxa, plus a few additional members of the dry forest ecosystem, occur in the archaeobotanical assemblage from Tibes (see below). Thus, assuming generally similar conditions about the time the site was occupied as exist today, it is reasonable to infer rather extensive coverage of subtropical dry forest in the Tibes landscape when the site was active, and one that almost certainly was more or less deciduous in aspect, depending on the prevailing climate conditions, as described above. It is worth noting also, considering the zooarchaeology, that the subtropical dry forest ecosystem has greater diversity of bird species compared with the wetter life zones on Puerto Rico (Ewel and Whitmore 1973:10–16; Murphy et al. 1995:186), with well-developed plant–bird dispersal mutualisms. The abundance of both birds and native edible fruits was probably attractive to humans settling in the area, and indeed both are part of the archaeological record at Tibes. Tibes is situated in an ecotone between the low coastal plain and the piedmont, on alluvial terraces fringed with riparian forests along the river. It is possible that the site was positioned and established in this relatively diverse setting to take advantage of particular resources associated with the different environmental zones: coast, coastal plain, and piedmont and lower cordillera forests, as well as freshwater aquatic and wetland habitats. If the site’s occupants were cultivators, a reasonable assumption based on the information presented above, this location may have helped to buffer some of the inherent constraints on agriculture in such a setting of relatively low rainfall and desiccating soil conditions. Scudder (1999:12) indicates that arable soils are located within a 2-km radius of Tibes,

Paleoethnobotanical Research at Tibes / 87 Table 5.1. Primary arboreal taxa found in the southern dry limestone region of Puerto Rico Indicator Tree Species Associated with the Subtropical Dry Forest Life Zone Acacia spp. (Fabaceae-Mimosoideae), acacia Bucida buceras (Combretaceae), úcar Bursera simaruba (Burseraceae), almácigo Capparis spp. (Capparaceae), burro, caper tree Cephalocereus royenii (syn. Pilosocereus royenii; Cactaceae), sebucán Guaiacum officinale (Zygophyllaceae), guayacán Guaiacum sanctum (Zygophyllaceae), guayacán blanco Leucaena glauca (Fabaceae-Mimosoideae), zarcilla Melicoccus bijugatus (Sapindaceae), quenepa Pictetia aculeata (tachuelo, fustic, Fabaceae-Papiloinoideae) Key Arboreal Taxa (Importance Values [IV] >10.0) for the Deciduous Forest Association Bursera simaruba, Pictetia aculeata, and Cephalocereus royenii, as above noted Coccoloba spp., e.g., C. krugii, C. microstachya (Polygonaceae), uverillo, uvillo Erithalis fruitcosa (Rubiaceae), black torch, jayajabico, tea Exostema caribaeum (Rubiaceae), yellow torch, West Indian quinine bark Gymnanthes lucida (Euphorbiaceae), crabwood, ramón, tabaco, yaití Pisonia albida (Nyctaginaceae), corcho Thouinia striata var. portoricensis (= T. portoricensis) (Sapindaceae), ceboruquillo

though they require management to counteract erosion and nutrient depletion (and see Murphy et al. 1995).

Synopsis of Paleoethnobotanical Research The paleoethnobotanical research reported here results from our excavations conducted during the period 1995–2003 in 11 locations across the site. The several stone-outlined monuments variously interpreted as ball courts (bateys) or plazas belong to the latest phase of occupation at Tibes (Elenoid culture series) (Curet et al. 2006; González Colón 1984), as do most of the deposits we tested and sampled. The paleoethnobotanical assemblage thus derives primarily from the later period of occupation, i.e., from deposits attributed to the Ostionoid (Elenan/Ostionan [Rouse 1992]) phase (ca. a.d. 800–1250). In a few locations strata representative of the earlier Saladoid component were encountered (Curet et al. 2006; Newsom and Curet 2000, 2003a, 2003b), including 1996 Units 2 and 3 and 2003 Unit N276 E105, all of which were located in the northern and northeastern sectors of the site (see Figure 3.1). In addition, the 1996 Unit 6 contained Saladoid ceramics in deeper levels, with Ostionoid materials in stratigraphically superior position. At

88 / Lee A. Newsom least two activity areas, with floor deposits and sets of postholes and/or cookingrelated features, and approximately eight midden deposits were sampled for plant remains. The 1995 Unit 1 in particular penetrated a dense shell midden deposit extending from level 3 to level 6. None of the current samples derive from within the bateys or from any definitive mortuary contexts. All together, 231 samples, including a total of 677 liters of bulk sediment sample, have been analyzed. This assemblage derives from 99 separate proveniences, with emphasis on floors and other contexts representative of householdrelated activities, e.g., cooking-area locus 1 in Operation 19 (OP19), a wellpreserved floor with an in situ ceramic griddle (buren) and grinding stones, thus equipment suggestive of food preparation. Further details on the proveniences sampled, taphonomic concerns, and sampling issues, as well as the field processing procedures, laboratory methods, and analytical procedures employed in the analysis can be obtained in Newsom (2007, 2009).

Results of Analysis Basic comparisons between the primary categories of plant material contained within the respective samples and proveniences analyzed from the site provide a basis by which to begin to compare the separate areas of the site. Archaeological seed remains (carbonized and mineralized categories) in general are very few (68 total), which is fairly typical of Caribbean sites.2 Carbonized wood is much more abundant. The dense midden sampled by Unit 1 produced the greatest quantities of carbonized wood (68.8 g total; 113 NISP [number of individual specimens]) for all areas tested, culminating also in the greatest number of identified specimens (111 individual fragments assigned taxonomically). This may be explained by the simple nature of the deposit as a collection zone for discarded fuelwood and other materials. Other locations that produced relatively large quantities of wood charcoal include OP19C and OP19D, due to concentrations of charcoal, ash, and fauna, and Macroblock Unit N184 E55 and Macroblock Unit N185 E56, the former including a well-preserved post feature (Feature 03-2) with large quantities of wood from the original post (160.30 g),3 and the latter having a high dispersion of pulverized wood (hence the low numbers identified [MNI] for this unit and adjacent units), an observation to which I return later. It is revealing to compare the stratigraphic distribution of wood charcoal from the excavations. Unit 1 wood residues were most concentrated in levels 3 and 4 (20–40 cm below surface; 57.2 g vs. 2.0) and underrepresentation (adjusted residual 4.0), are associated with three of the volcaniclastics: volcanic breccia, green volcanic breccia, and tuffaceous lithic sandstone. In contrast, the andesite porphyry, green tuff, and calcareous sandstone lithologies show no significant divergence in any structure from the overall mix (this despite the lack of any calcareous sandstone boulders in the two smallest structures surveyed, areas 5 and 9). Overall, boulder intermediate dimension values range from 4 to 109 cm, with

Quartz diorite Gabbro Andesite porphyry Volcanic breccia Green volcanic breccia Tuff Black tuff Brown tuff Green tuff Banded tuff Tuffaceous lithic sandstone Calcareous sandstone Limestone Packed biosparite Court average boulder size (intermediate dimension) and size range, cm

Boulder Lithology

–5.0 +2.9

+8.1 –5.7

–2.6

19.7 12–38

–3.0 +3.7 +2.5 +3.5 23.8 6–68

–2.0 22.4 7–105

+3.9

Area 5 Batey de una Hilera (n = 30)

–4.3

+2.0

+2.2 –4.3

Area 3 Batey del Cemí (n = 1,671)

–4.5 +2.5

Area 2 Batey de Herradura (n = 806)

29.4 10–109

+2.1

–2.9

–3.2

Area 6 Plaza Principal (n = 1,881)

28.0 4–80

+2.4

Area 7 Plaza de Estrella (n = 652)

25.2 10–50

–2.3 +2.0

+2.3

Area 8 Batey del Murciélago (n = 332)

27.9 7–50

–2.7

+2.8

Area 9 Batey del Cacique (n = 111)

Significant Differences (denoted by adjusted residual values), Each Pavement Area vs. All Areas

Table 8.2. Significant differences in mix of boulder rock types among surveyed pavement areas at Tibes, as determined by chi-square analysis

Boulder Lithology Survey at the Tibes Ceremonial Site / 183 a mean size of approximately 25 cm. Mean values in individual structures range from 19.7 to 29.4 cm (Table 8.2).

Discussion The overall match in the range of rock types between Tibes pavements and river deposits strongly supports the interpretation that the Elenan Ostionoid builders used the Portugués River bed, banks, and floodplain as their routine material sources for construction of the ball courts and plazas. There are two possible types of explanation for the significant quantitative differences in lithologic mix between the sampled riverbed sites and the group of pavements, and among the individual ball courts and plazas. One type of explanation relates to natural variation in river-transported sediment through time, and the other relates to a nonrandom selection of building material by the indigenous people. If the percent lithologic mix of boulders being carried downstream by the Portugués River were to significantly shift through time, this could account for the overall differences between structures built more than 800 years ago and the riverbed today. If the same variation was significant on the order of decades, it could also possibly account for the differences among structures, assuming that none were built simultaneously. In coming years it will be possible to directly test for fluctuation in the Portugués River bed mix of boulder rock types by repeated sampling. Such fluctuation could result from shallow landslides that act as primary agents for downslope sediment delivery to streams in Puerto Rico (Divakarla and Macari 1998; Larsen and Torres-Sanchez 1998). Local, major landslides delivering pulses of boulder material to the stream from discrete bedrock source areas could provide a reasonable natural mechanism for a variation of dominant lithologies moving downstream. An evaluation of the magnitude and frequency of such catastrophic mass-movement processes in this particular drainage basin would give insight on the potential for decade- and century-scale variations in the lithology mix of channel sediment load. Alternatively, the mix of rock types in the Portugués River bed and banks may have remained essentially unchanged through the present, but selective use was made of the available array of construction materials. The builders of particular courts and plazas may have established explicit (although not absolute) preferences for some lithologies over others, based on their appearances fresh from the riverbed. If this was the case, then the evidence suggests that the overall preference varied over time or that the use of materials with distinctive symbolic meaning was altered in accordance with the specific function of each structure. Physical factors correlating with lithology may have also played roles in selective use. For workers filling spaces in partially laid pavements, boulder size or shape may have guided selection. A particular preferred boulder size may have been a guiding factor in the

184 / Rice-Snow, Castor, Castor, Grigsby, Fluegeman, and Curet construction of each court or plaza, as suggested by the differing mean boulder sizes for pavements in Table 8.2. Some rock types of undesirable surface texture or higher density (L. Weidman, personal communication 2002) may have been avoided by experienced transporters of boulders from the riverbed. Cobbles of some lithologies may have been set aside as raw material for the making of household items or for other community use (J. Walker, personal communication 2006). The data collected for this study do provide some basis for evaluation of boulder size as a primary factor in selective use of available lithologies. The different lithologies in the pavements have different mean boulder sizes, ranging from 22.1 to 28.7 cm (Table 8.3). Therefore, a search for boulders of a particular size from the riverbed would have the potential to overrepresent some rock types and underrepresent others. In general, we would expect the lithologies with larger mean boulder size (Table 8.3) to be overrepresented in pavements showing large mean boulder sizes (Table 8.2). The inverse would be true for lithologies and pavements characterized by small mean boulder sizes. None of these predictions bear out with any consistency as we examine the actual overrepresentations and underrepresentations of lithologies in different pavements. There is no evidence here that boulder size and lithology associations can help explain the patterns seen. One particular result is notable in the boulder lithology rankings by size (Table 8.3): the calcareous sandstone boulders selected for the pavements have the smallest mean size of any rock type, quite possibly related to the need to transport them by hand over a longer distance to the site.

Calcareous Sandstone Boulders All of the lithologies represented in the structures at Tibes are found in the nearby Portugués River bed except one, a friable calcareous sandstone. This lithology is represented in the surveyed population at Tibes by 52 angular slabs ranging from 12 to 41 cm intermediate dimension. Several of these sandstone boulders display petroglyphs, which are rare features overall in the park. These calcareous sandstone clasts also lack the distinctive rounding apparent in other boulders making up the structures, further supporting our interpretation that they were not delivered to Tibes by fluvial transport. A review of U.S. Geological Survey geologic quadrangle coverage for the area indicates no calcareous sandstone outcrop areas within the Portugués River drainage basin leading to Tibes. No channelized flow or other natural, gravity-driven transport process could have delivered the sandstone boulders to the site: they must have been hand carried from a local or remote source. The nearest calcareous sandstone outcrops outside the basin are found within a kilometer of Tibes on the adjacent coastal plain. We have carried out a preliminary sampling study of cal-

Boulder Lithology Survey at the Tibes Ceremonial Site / 185 Table 8.3. Average boulder sizes for the rock types represented in the surveyed Tibes pavements Boulder Lithology, Ranked by Average Boulder Size (descending) Green volcanic breccia Quartz diorite Banded tuff Volcanic breccia Andesite porphyry Limestone Gabbro Black tuff Packed biosparite Tuffaceous lithic sandstone Green tuff Brown tuff Tuff Calcareous sandstone

Average Boulder Size (intermediate dimension, in cm)

Boulder Size Range (cm)

No. of Boulders

28.7 28.5 27.9 27.6 27.6 27.0 26.6 25.8 25.2 25.1 24.8 24.2 23.2 22.1

11–80 10–105 13–80 10–109 7–79 11–60 7–90 10–74 12–45 6–105 4–64 10–70 10–70 12–41

181 713 67 932 330 36 485 79 65 981 690 507 419 52

careous sandstone boulders at Tibes and selected nearby sites to evaluate the possibility that this material was brought in from a single, local source.

Methods Once approval was granted for limited sampling within the Tibes site, we collected calcareous sandstone samples from two boulders (Figure 8.1) with the oversight of José Lugo, Park Director. The sample designated Tibes #1 was obtained from a boulder near the southeast corner of pavement area 6, Plaza Principal. Sample Tibes #2 was taken from a boulder near the northern limit of the eastern boundary wall of pavement area 3, Batey del Cemí. In order to avoid defacing the aboveground surfaces of these partly buried boulders, the workers temporarily removed them and took approximately fist-sized samples from their undersides with a rock saw. We collected material from three likely sites outside the drainage basin near Tibes (Figure 8.2), designating samples from two or more sandstone boulders at each site for further analysis. In all cases, the samples were taken from boulders lying loose on the ground surface. The Cemetery Site South is about 900 m southwest of Tibes, next to a cemetery that itself is just west of the housing development Barriada Jaime L. Drew. The rock bed is exposed at the top of a hill, and the

Figure 8.1. Locations of two boulders sampled from the structures at Tibes. Italic numbers identify pavement areas listed on Table 8.2.

Boulder Lithology Survey at the Tibes Ceremonial Site / 187

Figure 8.2. Locations of three possible calcareous sandstone source areas, external to the Tibes site, included in the preliminary sampling program.

samples were collected from the base of the hill. The Cemetery Site North is adjacent, about 200 m to the northwest of the first site, on the opposite side of an east–west road. The boulders appeared to be derived from the same hill as on the first site, although a larger hill rising to the north also is partly based on sandstone (Juana Diaz Formation) bedrock. The third location, the Ponce Gardens Site, is about 2,700 m east of Tibes on a promontory at the southern edge of a terrace occupied by the Jardines de Ponce housing development. While it is common geologic practice to look for rock samples “in place” (still attached to bedrock), it is reasonable to believe that the Pre-Taíno builders of the Tibes structures would readily exploit any site offering the desired building material and would likely prefer sources of appropriate-size, loose boulders that would not require quarrying. Interestingly, many of the boulders sampled at the above sites were relatively flat slabs, similar in shape to the ones found in the Tibes structures. The samples were thin-sectioned and microscopically studied for sedimentary rock characteristics and microfossil content. Key characteristics used to mark differences and similarities between samples were the sizes of sedimentary grains, degree of sorting (small vs. large variety of grain size in a sample), composition of the grains (biologically formed particles vs. detrital quartz grains and volcanic fragments), microfossil species, and condition of remains. These led to further interpretations of the depositional environment of the sediments and their approximate age.

188 / Rice-Snow, Castor, Castor, Grigsby, Fluegeman, and Curet

Results The two Tibes samples are not similar and appear to have been brought from different source locations. Tibes #1 is very carbonate-rich sandstone including approximately 20 percent detrital fragments, is well sorted, and lacks any specific age-indicator fossils. Tibes #2 is properly classified as a limestone, with approximately a 5 percent detrital component, and shows significant differences from all other samples taken in this study. It is very poorly sorted and includes types of foraminifera microfossils that do not appear in the other thin sections. The sediments forming this rock appear to have been deposited in a fore-reef platform environment, on the deep offshore side of a reef. The estimated age for Tibes #2 is Cretaceous. The rock samples obtained from the three sites near Tibes are all carbonaterich sandstones, early Eocene in age. The different boulder samples from Cemetery Site South match up well, showing little difference in composition, with a 16 to 26 percent detrital component. Rocks from Cemetery Site North contain approximately a 32 percent detrital component, but the two boulders sampled do not match up as well in contained fossil species and degree of sorting. Samples taken from the Ponce Gardens Site are 21 to 34 percent detrital, with some variation in grain size. The samples from all three sites lack planktonic foraminifera microfossils, suggesting that the sediments were deposited in a shallow nearshore environment.

Discussion It is possible that the Tibes #1 sample from pavement area 6 could have been obtained from any of the three examined sites near Tibes. The closest match is one of the rock samples from Cemetery Site North. However, Tibes #1 has better sorting (constancy of grain size) than any other sample examined and lacks green algae fragments, which are found in most samples from the three possible source sites. In these ways it represents an extreme rather than a typical case, when compared to the samples from the three sites. Tibes #2, obtained from pavement area 3, is unique in the suite of samples, showing a strong mismatch with the rocks from examined areas near Tibes. The two sampled boulders in the Tibes structures were brought from different source sites by the Pre-Taíno community. More comprehensive sampling will be needed to determine whether additional source sites are represented among the other 50 Tibes boulders classified as calcareous sandstone. Further work will also help illuminate whether the variation in source area is one mainly between the different structures or also among boulders within each of the structures. There are a number of reasons that people may expend labor to bring rock of a

Boulder Lithology Survey at the Tibes Ceremonial Site / 189 distinctive type to a site with abundant stone resources. For example, the imported rock may have better material properties for a specific use (Brink and Dawe 2003), it may serve a decorative function (Miriello and Crisci 2006), or it may have symbolic societal meaning in connection with its geographic point of origin (Ogburn 2004). Any of these three reasons has the potential to apply to the presence of exotic calcareous sandstone blocks at Tibes. The friability of the sandstone and surviving petroglyphs on some clasts suggest that it was a preferred rock-carving material. While in the case discussed by Ogburn (2004) building stone was exported from the core of an empire to its periphery, at Tibes there may have been symbolic political significance in the convergence of exotic stones from the periphery on the central ceremonial site.

Conclusions We conclude, first, that the likely source area for most of the boulders used in the construction of the ball courts and plazas at Tibes is the nearby Portugués River. All but one of the lithologies present in the site are represented in the riverbed. Second, overall percentages of some lithologies vary significantly between the river source area and the site structures, and they also vary among those structures. This suggests natural variation of the percent mix of lithologies in the riverbed through time, selective use of boulder lithologies by the builders of the Tibes structures, or both factors in combination. Third, the calcareous sandstone boulders within the structures at Tibes were not derived from the riverbed, having been transported by hand from sources outside the river drainage basin. Fourth, the calcareous sandstone boulders were brought to the Tibes site from two or more distinct points of origin. Evaluation of this aspect of the Tibes structures warrants further focused work.

Acknowledgments We very much appreciate on-site assistance provided by José Lugo, Carmen Martínez Ajá, and other staff of the Centro Ceremonial Indígena de Tibes, and access to the site approved by the city of Ponce. Additional support in Puerto Rico was provided by Lee Newsom (Penn State University) and Ray Petty (Inter American University of Puerto Rico). Harlan Roepke and Theresa Boundy of Ball State University provided petrographic advice in the early stages of the project. We thank an anonymous reviewer for advice improving the manuscript. Jill Seagard (Scientific Illustrator, Field Museum Department of Anthropology) produced Figures 8.1 and 8.2 for this report. Funding for this study was provided by the national

190 / Rice-Snow, Castor, Castor, Grigsby, Fluegeman, and Curet Sigma Xi Scientific Research Society, the North Central Section of the Geological Society of America, the Indiana Academy of Science, the Ball State University Office of Academic Research and Sponsored Programs, the Ball State Department of Geology, and National Science Foundation grant #BCS-0106520 (L. A. Curet, PI).

9 Ancient Bones Tell Stories Osteobiography of Human Remains from Tibes Edwin F. Crespo-Torres

Between 1975 and 1982, a large number of human burials were discovered during archaeological excavations at the prehistoric ceremonial center of Tibes. Thirteen years later a systematic skeletal analysis was performed on these remains. General information concerning age at death, sex, height, and some pathological conditions was obtained from 126 individuals. This information was used to develop an osteobiographical profile of the people who were born, lived, and died at Tibes. This chapter summarizes the results obtained for Tibes and compares them to the information obtained from other archaeological sites from which human remains have been recovered in Puerto Rico, including Punta Candelero (n = 85) and Paso del Indio (n = 152) (Crespo-Torres 2000a) and Maisabel (n = 34) (Budinoff 1991) (Figure 9.1).

Previous Reports on Human Remains from the Ponce Region Froelich Rainey (1940) was the first archaeologist to report human burials from the Ponce region. In his report he describes 24 sets of human remains he excavated at the site of Canas during the 1930s. Regrettably, the current location of these remains is unknown and they may possibly be included in the hundreds of skeletal remains from Puerto Rico that have lain hidden and unstudied in storage in various American museums (Drew 2005; El-Najjar 1977) since as early as 1918 (Aitken 1918). Forty-five years after Rainey’s work, the discovery of human remains in the area of Ponce was reported at the sites of El Bronce and Caracoles. Ten individuals were recovered at El Bronce (McClung and Hughes 1985), and nine burials were

Figure 9.1. Map of Puerto Rico showing the locations of archaeological sites where human remains have been recovered.

Ancient Bones Tell Stories / 193 reported from Caracoles (González and Rodríguez 1985). Archaeological context and the osteological analysis of the remains from El Bronce were included in the final report of the project. However, no information for the remains from Caracoles has been published or is available from other sources. During the early 1990s another project at the site of Caracoles, initiated due to the construction of the Plaza del Caribe Shopping Mall, led to an archaeological data-recovery operation. The final report for this project contains an osteological analysis I conducted for five human burials recovered during these excavations (Molina and López 1995). Finally, excavations at the significant archaeological site of Maruca between 1994 and 1995 also produced human remains. The site’s importance is based on its antiquity, as an absolute date of 4950 b.p. was obtained. Such dating makes the 11 human burials recovered from the site the oldest in Puerto Rico. The final report includes an osteological analysis conducted by me in 1997 (Crespo-Torres 2004).

Human Remains from the Ceremonial Center of Tibes Archaeological excavations were undertaken at Tibes between 1975 and 1982 by an avocational archaeological group under the direction of Juan González Colón (Alvarado Zayas 1981; González Colón 1984; see also Alvarado Zayas and Curet, this volume). According to González Colón (1984), an unspecified “large number” of human burials were discovered. Unfortunately, no field data or records have been published that will allow determination of any diachronic and synchronic provenience for these burials. In fact, the only information available on this matter consists of very succinct data included in González Colón’s (1984) Master’s thesis, indicating the discovery of not fewer than 39 human burials in midden I and plazas 3, 6, 7, and 8. A similar situation is encountered in the Master’s thesis of Alvarado Zayas, in which a brief and incomplete description prepared by Fernando Luna Calderón of 57 human remains unearthed from the Plaza Principal (Structure 6) is included (Alvarado Zayas 1981:32–42). Using these data I was able to determine only the general distribution of these burials, but not their specific synchronic and diachronic context, and much less how they relate to each other. Unfortunately, the burials were not awarded the attention they truly deserved. The absence of a final report on the excavations has, so far, prevented not only the thorough analysis of funerary behaviors but also a complete bioarchaeological study.

Methods of Skeletal Analysis It must be said that most of the skeletal material was in a poor state of preservation, in part due to natural factors such as soil acidity and fragmentation caused by

194 / Edwin F. Crespo-Torres taphonomic process and in part due to poor storage conditions. Thirty-eight bags full of human bones were found stacked upon each other within the same box, which caused the postmortem fragmentation of many bones. In addition, each bag contained intermixed bones for one or multiple individuals. During the analysis, standard procedures were used in the reconstruction, inventory, and analysis of human skeletal material (Buikstra and Ubelaker 1994). A total of 126 individuals were included in this analysis (Crespo-Torres 1998). Each skeleton was systematically examined to obtain a biological profile. Sex, age at death, and height were obtained by the application of morphological and quantitative methods, as outlined in such sources as Bass (1987), Krogman and Iscan (1986), Scheuer and Black (2000), Schwartz (1995), Steel and Bramblett (1988), Ubelaker (1989), and White and Folkens (2005). Evaluation of skeletal pathology followed procedures and categories outlined in Aufderheide and Rodríguez-Martín (1998), Mann and Murphy (1990), Ortner (2003), and Steinbock (1976). After the analysis, the remains were stored in acid-free boxes.

Results Distribution of Age at Death and Sex The sample from the site of Tibes was composed of 31 adult females (25.0 percent), 25 adult males (20.0 percent), 62 adults whose sex could not be estimated because of the poor state of preservation (49.0 percent), and 8 subadults (6.0 percent) (Table 9.1). After adjusting the age-at-death distribution, a total of 95 of these burials were used in a recent paleodemographic study conducted by Antonio Curet (2005).

Height Estimation Unfortunately, the poor state of preservation and the fragmentation of the remains did not allow the determination of height for most of the individuals. Height estimates were possible for only two of the individuals recovered at Tibes. Estimates were calculated of 157 cm (5′1″) for an adult female and 160 cm (5′3″) for an adult male (Table 9.2).

Pathological Conditions Various pathological conditions were detected in both teeth and bones, which offers a panorama of what kinds of diseases were present in the ancient population of Tibes. Table 9.3 summarizes the pathological conditions that are present in Tibes individuals. Dental Pathologies. The sample exhibits three main dental pathological conditions: caries (21.0 percent), dental wear (13.4 percent), and calculus deposition (10.3 percent). Dental caries has been the most common dental disease among hu-

Ancient Bones Tell Stories / 195 Table 9.1. Sex/age distribution in Tibes Age/Sex Adult Male Adult Female Subadult Indeterminate Total

n

%

25 31 8 62 126

20.0 25.0 6.0 49.0 100.0

Table 9.2. Stature estimation at Tibes and other archaeological sites in Puerto Rico Site

Age/Sex

Stature

Tibesa

Adult male Adult female

160 cm (5′3″) 157 cm (5′1″)

Punta Candelerob

Adult male Adult female

158 cm (5′2″) 150 cm (4′11″)

Paso del Indioc

Adult male Adult female

159 cm (5′2″) 150 cm (4′11″)

Maisabeld

Adult male Adult female

159 cm (5′2″) 145 cm (4′9″)

a

c

b

d

Crespo-Torres 1998. Crespo-Torres 1991.

Crespo-Torres 2001a. Budinoff 1991.

mans since the origins of agriculture (Cohen and Armelagos 1984). Caries is initiated by demineralization of the enamel by organic acids that are locally produced by bacteria and is characterized by decalcification and disintegration of the hard dental tissues. In addition to causing demineralization, the bacteria also destroy the protein content of teeth (dentine). Bacteria and dietary carbohydrate diet are important factors in the initiation of enamel caries. However, the cariogenic process can be the result of multiple factors within the oral environment (Lazzari 1976; Ortner 2003; Spouge 1973). Twenty-six out of 126 individuals (21.0 percent) showed caries lesions. Dental wear is defined by some authors as the erosion process of the coronal enamel. It is not itself a pathological condition but rather the natural result of masticatory stress upon the dentition in the course of both alimentary and technological activities. In general terms, wear includes both attrition and abrasion (Bhaskar 1973; Gorlin and Goldman 1970; Guita 1984; Ortner 2003; Spouge

196 / Edwin F. Crespo-Torres Table 9.3. Pathological conditions observed at Tibes and other archaeological sites in Puerto Rico Site Tibes (n = 126)

Punta Candeleroa Paso del Indiob,c (n = 85) (n = 129)

Pathology

n

%

n

%

n

%

Maisabeld (n = 34)

Caries Dental wear LSAMAT Calculus Periodontitis Periapical abscess AMTL Periostitis Osteoarthritis Benign tumors Trauma Porotic hyperostosis

26 17 1 13 3 2 3 6 5 2 2 1

21.0 13.4 0.8 10.3 2.4 1.6 2.4 4.8 3.9 1.6 1.6 0.8

25 41 10 24 20 11 15 8 7 — 1 —

29.4 48.2 11.8 28.2 24.0 13.0 18.0 9.4 8.2 — 1.2 —

45 25 11 17 11 18 30 38 17 2 9 13

35.0 19.4 11.4 13.1 9.0 14.0 23.3 29.5 13.2 2.0 7.0 10.1

Present Present 1 Present Present Present Present 2 Present — 1 —

Note: LSAMAT = lingual surface attrition of maxillary anterior teeth; AMTL = antemortem tooth loss. a Crespo-Torres 1994. b Crespo-Torres 2000a. c Crespo-Torres 2001a. d Budinoff 1991. Budinoff reports these conditions as present, but the frequency in the sample is not shown.

1973). Seventeen out of 126 individuals (13.4 percent) studied from Tibes show evidence of dental wear. On the other hand, a particular wear pattern known as lingual surface attrition of maxillary anterior teeth (LSAMAT), first reported by Turner and Machado (1983) and Irish and Turner (1987), is also present in one female skeleton from Tibes. The cited authors argue that the use of the upper anterior teeth and tongue to manipulate a gritty cariogenic food such as manioc roots may have produced this unusual wear related to caries. Dental calculus is the mineralization of bacterial plaque. The exact etiology of calculus remains unknown. However, some dental specialists have suggested that higher amounts of calculus deposition correlate with a dietary imbalance characterized by high carbohydrate and low protein intake (Lazzari 1976; Stanton 1969). Thirteen out of 126 individuals (10.3 percent) show evidence of dental calculus at Tibes.

Ancient Bones Tell Stories / 197 Other dental pathological conditions such as periodontitis, periapical abscess, and antemortem tooth loss are also present in some individuals from Tibes. Unfortunately, as mentioned above, the bad state of preservation of the skeletal sample does not allow estimating more accurately the real prevalence of these pathological conditions among the prehistoric inhabitants of Tibes. Finally, although not a pathological condition but rather a dental morphological trait of anthropological importance, shovel-shaped incisors were detected in some individuals from Tibes. Unfortunately, heavy dental wear present in the dentition of Tibes inhabitants limited the morphological dental traits analysis. Bone Pathologies. Other pathological conditions that affected bony tissue are also present in the skeletal samples from Tibes. Many of the lesions present in bones caused by infections are produced by microorganisms such as streptococcus, staphylococcus, and gram-negative bacteria that disseminate as part of a systemic condition or as a consequence of untreated fractures. Analysis of the material discloses conditions known as osteitis or bone inflammation, which is an immunological response to an infection. When the inflammation is circumscribed to the external surface of a bone, it is denominated as periostitis, but if the infection affects the inside of the bone shaft it generates a much more severe condition known as acute osteomyelitis (Ortner 2003). At least six individuals (4.8 percent) from the Tibes sample showed clear evidence of periostitis in their bones. Another condition present in five individuals from Tibes is the osteoarticular lesion. The most common cause of this condition is arthritis, associated with a normal degenerative process during old age, physical activities, and trauma. Despite the general condition of bad preservation of the skeletal sample, the analysis detected five cases (3.9 percent) in which this kind of lesion was present in both finger bones and vertebrae. Two kinds of benign bone tumors were detected at Tibes. An adult female showed a small button osteoma in the frontal bone and an adult male presented another benign tumor of the ear canal known as auditory exostosis. The etiology of this latter lesion can range from genetics to some activities such as continuous diving in deep, cool water (Aufderheide and Rodríguez-Martín 1998:254–256; Frayer 1988; Kennedy 1989). A hematological disorder known as porotic hyperostosis is present in one adult male. There is much debate over the etiology and proper classification of this condition. However, most researchers associate this condition with iron-deficiency anemia. Two cases of antemortem bone fracture were detected. One adult male had an antemortem fracture in the right ulna. Another adult male showed an antemortem fracture in the left femur. In both cases, the callus developed during the bone healing process is visible.

198 / Edwin F. Crespo-Torres

Cultural Modification One cultural modification that can be found in the skeletons of the indigenous populations of Puerto Rico is the practice of artificial cranial deformation. This consists of the artificial alteration of the normal contour of the skull by applying an external force. This alteration was identified in three adult males, who presented the type known as fronto-occipital deformation (Figure 9.2a). It must be clarified, however, that we do not discard the possibility that this practice may have been present in other individuals from Tibes, but it cannot be detected because of the bad state of preservation of the remains.

Discussion The results obtained from Tibes are compared with those from three other archaeological sites at which a number of human remains were found (Table 9.3). These three sites are Punta Candelero, Paso del Indio, and Maisabel (Budinoff 1991; Crespo-Torres 1994, 2000a, 2001a). Unfortunately, we do not have the frequency of each of the conditions present in the sample of 34 individuals recovered from the Maisabel site. Punta Candelero is an early Ceramic site located on the eastern coastal plain of the island, less than 300 m from the shore. The site occupies a good portion of a sandy peninsula that extends about 1 km into the Caribbean Sea. The archaeological deposits are composed of two main assemblages or occupations. The first one belongs to the La Hueca Complex, while the second one belongs to the Cuevas style (a.d. 400–600) (Rodríguez 1989, 1991). A total of 85 human remains were found at the site and belong to the Cuevas style (Crespo-Torres 2000a). The site of Paso del Indio is located on the north-central coast of Puerto Rico on the floodplain of the Cibuco River in the boundary between the north coastal plain and the karst zone of the island, approximately 6.6 km from the shore. This site seems to have been occupied from the Archaic to Chican Ostionoid (a.d. 1200–1500) times (Walker 2005). A total of 152 burials found at Paso del Indio are from the Elenan Ostionoid component (Crespo-Torres 2000a). Maisabel, a Saladoid-Ostionoid site, is located on the shoreline of the Atlantic Ocean in the lowlands of Puerto Rico’s north-central coastal plain. A total of 34 burials were excavated from this site and their dates range from the Early Saladoid (Hacienda Grande style) to Ostionoid periods (Siegel 1992). An osteological study was performed by Linda C. Budinoff (1991); demographic profiles, height estimations, health conditions, and cultural modifications were determined for all 34 individuals. A paleodemographic study done by Antonio Curet (2005) made a comparative sex/age profile among Tibes, Punta Candelero, and Paso del Indio. The three

Ancient Bones Tell Stories / 199

Figure 9.2. Artificial frontooccipital cranial deformation: (a) skull from Tibes, with arrows showing the frontal and occipital compression (photo taken by E. Crespo-Torres); (b) a Huecoid biomorph amulet holding in its claws an intentionally deformed human head—note the straight line from the nose tip to the forehead (dark line) (Center of Archaeological Research, University of Puerto Rico, photo taken by E. Crespo-Torres).

samples, according to Curet (2005:212), showed marked differences in their sex distribution. The population from Punta Candelero tended to have a higher number of adult males, while the one from Tibes was the opposite. In the case of Paso del Indio, a “normal” distribution was reported, with all sexed adult skeletons equally divided between the two sexes. Curet has argued that these differences seem to have been produced by multiple processes and no single explanation can account for all of the differences. Processes related to the migrant nature of the population and some form of bias in the sample related to the poor condition of the skeletal remains could be responsible for the sex distribution differences among the three skeletal samples (Curet 2005:212–213). On the other hand, the age-at-death distribution for the three samples does not match any ideal model (Curet 2005:213–214). Age-at-death distributions in ideal paleodemographic models for preindustrialized populations are characterized by a large number of individuals between the ages of 0 and 5 years, then the number

200 / Edwin F. Crespo-Torres of individuals drops rapidly after that. A second rise is, then, observed at the older ages of 50 and over (Márquez and Hernández 2001). The distribution from the skeletal remains from Paso del Indio is the only one that gets close to this ideal pattern by showing a sharp peak of individuals in early age between 0 and 10 years old, but without the final rise at the older ages. Curet has argued that a possible explanation for the “abnormal” distribution of the Punta Candelero sample is that it may represent the migrant nature of the Candelero populations that tended to be dominated by young and middle-age males with fewer subadult and adult females and children. On the other hand, the difference observed in the Tibes population is the result of the poor state of preservation that probably affected the skeletal remains of the subadult component (Curet 2005:214). Age-at-death distributions for Tibes, Punta Candelero, and Paso del Indio are shown in Table 9.4. The sex/age distribution in the population from the Maisabel site apparently showed a 1:1 sex ratio in which males lived longer than females. According to Budinoff (1991:122), this suggests a low level of warfare and that many females were dying during the childbearing years. Likewise, few people of either sex lived well into old age. However, Budinoff emphasizes that the sample is too small to make any major demographic inference, although sufficient for a preliminary one. Heights in the Tibes sample are similar to the height indexes obtained in other pre-Columbian skeletal samples in Puerto Rico (see Table 9.2) (Budinoff 1991; Crespo-Torres 1991, 1998, 2001a). In addition, some stress indicators present in teeth and bones (e.g., enamel hypoplasia, porotic hyperostosis, and periostitis) reveal the strong relationship between growth suppression in childhood and attainment of adult body size, including terminal height. The close ties between stress, especially poor nutrition, and stature are abundantly documented in research developing out of a growing interest in anthropometric history (Larsen 1997). Therefore, future studies combining both height estimation and the frequencies of stress indicators could not only provide a realistic profile of height but also help identify those health conditions that may have affected the growth process for the entire pre-Columbian Caribbean population. There is no doubt that the dental and bone pathologies present in Tibes (Table 9.3), as well as the cultural modifications, are also found in other archaeological remains at sites in Puerto Rico, such as Punta Candelero, Paso del Indio (CrespoTorres 2000a), and Maisabel (Budinoff 1991). As mentioned above, a multiplicity of factors can act as etiological agents and contribute to the development of caries. However, the principal factor associated with the development of cariogenic lesions in the prehistoric agriculturalist groups in the Caribbean area is the consumption of resources with high concentrations of carbohydrates such as manioc (Manihot esculenta), maize (Zea mays), and probably some comestible fruits. The archaeological (González Colón 1984:262–265) and paleobotanical (Curet and Newsom 1998; deFrance and Newsom 2005; see Newsom, this volume) evidence

Ancient Bones Tell Stories / 201 Table 9.4. Age-at-death distribution observed at Tibes, Punta Candelero, and Paso del Indio archaeological sites Age

Tibesa

Punta Candelerob

Paso del Indiob

0–4 5–9 10–14 15–19 20–24 25–29 30–34 35–39 40–44 45–49 50–54 55–59 60+ Total

3.000 2.000 1.000 2.000 21.141 14.140 11.140 16.140 10.140 8.140 7.140 .000 .000 95.981

6 5 0 1 9 14 9 11 7 14 9 0 0 85

41 20 5 2 7 11 9 13 13 4 4 0 0 129

a

After adjusting age-at-death distribution (Curet 2005). Crespo-Torres 2000a.

b

from Tibes indicates that at least two of these food sources (manioc and comestible fruit) were present at the site and are probably associated with the development of caries in the Tibes skeletal sample. On the other hand, stable isotope analysis done in skeletal materials from Paso del Indio and Maisabel indicated differences in delta (d) values of the isotopes from both sites (Stokes 1998, 2005). According to this analysis it appears that some individuals from Paso del Indio obtained substantial amounts of carbohydrates from C4 plants (e.g., maize or other tropical grasses), whereas other people consumed more C3 plants, such as manioc and other staples. Contrarily, the results obtained at Maisabel show that C3 plants (e.g., manioc and other staples) were the major energy source there. At the present, a team composed of William Pestle, Antonio Curet, and myself (see Pestle, this volume) is conducting a major stable isotope analysis on the skeletal samples from Tibes and other osteological collections from different archaeological sites from Puerto Rico (including Punta Candelero and Paso del Indio). It is expected that the results from this analysis will help fill a gap in our knowledge of the diet of the ancient inhabitants of these sites. The dental wear present in the skeletal sample from Tibes is probably related to factors such as the type of hard food consumed, the presence of gritty or small particles of clay or sand adhered during the food preparation, and the use of teeth

202 / Edwin F. Crespo-Torres as tools. These factors probably contribute also to the high frequencies of dental wear present at the Punta Candelero, Paso del Indio, and Maisabel sites. Lingual surface attrition of maxillary anterior teeth (LSAMAT) is also present in the skeletal samples from Punta Candelero and Paso del Indio (Crespo-Torres 1994, 2000a). At these sites this particular dental wear is associated with caries and is reported only in females. According to some ethnohistorical accounts, the females in the indigenous groups in the Greater Antilles at the time of contact were responsible for food processing, especially of cassava (Sued-Badillo 1989). It is possible that at the two other archaeological sites mentioned above the difference in the presence of this kind of wear only in female skeletons might be related to a sexual division of labor. It is not really clear whether LSAMAT is present in the skeletal sample from Maisabel (Budinoff 1991). Nevertheless, the presence of this particular wear pattern but without caries is also reported in males (Budinoff 1991; Crespo-Torres 1994), probably related also to the use of teeth as a tool when making different artifacts (e.g., weapons, fishing nets, or fiber baskets). However, reanalysis of the Maisabel skeletal collection is necessary to clarify the presence or absence of this particular type of dental wear. In terms of the dental calculus accumulation present at Tibes, two factors may have contributed to this condition. First, the consumption of high carbohydrate products in the diet and, second, poor dental care habits among the group. This condition is also found in the skeletons from Punta Candelero, Paso del Indio, and Maisabel. According to Scott and Turner (1997), shovel-shaped incisors are present in 90 to 95 percent of the population for Amerindian groups, so its presence at Tibes is no surprise. The presence of shovel-shaped incisors is also reported in skeletal material from other archaeological sites in Puerto Rico (Crespo-Torres 1987, 1994, 2000b, 2001a, 2001b; McClung and Hughes 1985). No information about the presence or absence of this morphological dental trait in the Maisabel sample was reported. As mentioned before, heavy dental wear present in the dentition in most of the skeletal collections, not only from Puerto Rico but also in the rest of the Caribbean area, is at least one important factor that limited the observation and analysis of morphological dental traits. Most of the bone pathological conditions present in the skeletal samples from Tibes are also reported at the Punta Candelero, Paso del Indio, and Maisabel sites. In general terms, the relationship between the demographic profile and the pathological conditions suggests that the populations from Tibes, Punta Candelero, and Maisabel showed more stable health conditions than those at the Paso del Indio site. For example, periostitis is the most frequent bone pathological condition reported at Paso del Indio, with a high frequency in subadult skeletal samples.

Ancient Bones Tell Stories / 203 On the contrary, at Tibes, Punta Candelero, and Maisabel the frequency in the sample is lower and the condition occurs more often in adult individuals than at Paso del Indio. There is no doubt that the high frequency of periostitis and other pathological conditions, such as enamel hypoplasia, porotic hyperostosis, and cribra orbitalia, among the remains from Paso del Indio is an indication of a severe nutritional stress situation that affected specifically the subadult group in this population (Crespo-Torres 2000a:168–69). Geomorphological studies conducted at Paso del Indio confirm that periodic floods constituted the major depositional process at the site and that the frequency and intensity of floods varied during different periods of occupation, affecting residence and causing loss of property, severe disruption of village life, and interruptions to subsistence activities such as agriculture (Clark et al. 2003; Walker 2005), consequently leading to a nutritional stress condition that affected the population and in particular the subadult groups (Crespo-Torres 2000a:187–193). On the other hand, the auditory exostosis detected in one individual from Tibes is the first case reported in Puerto Rico. Frayer (1988) reported this condition in the human remains from the late Mesolithic site of Vlasac (6300–5300 b.c.) in Yugoslavia. In his data he found a much greater prevalence in male skeletons. Frayer notes the evidence for heavy reliance on fish at the site and speculates that the fish would probably have been caught using hand nets in a method that may have involved diving in cold water. If diving in cold water is the primary cause of this pathology, then it is surprising to find it at Tibes, which is 8 km inland. Contrarily, Punta Candelero and Maisabel are both located on the coast, but evidence of this condition is not reported from these sites. The etiology of this tumor remains unknown, but a possible candidate is a chronic, low-grade infection of the auditory canal, as suggested by Ortner (2003:517). The frequency of bone trauma at Tibes, Punta Candelero, Paso del Indio, and Maisabel is considerably small. In general terms, the trauma in the populations from these four sites consists of antemortem fractures probably related to physical activities. However, traumas related to interpersonal conflict are also reported for Tibes and Maisabel. At Tibes one adult male shows antemortem fracture in the right ulna midshaft. This kind of fracture is known as a parry fracture or defense fracture and results when a person raises his or her arm to prevent being struck in the face or head (Mann and Murphy 1990). According to Budinoff (1991:117–118), two adult males from Maisabel show probable evidence of a violent death. The first was supposedly killed with a stingray spine projectile thrust into the chest. The second individual shows a series of perimortem cut marks at the right humeral midshaft made with a bifacial tool. Nevertheless, these cases are not necessarily associated with interpersonal violence. The parallel position and location in the rib cage in which the stingray projectile was found during the excavation suggests that its position in relation to the

204 / Edwin F. Crespo-Torres body could be related to natural taphonomic deposition (Haglund and Sorg 1997, 2002) and not to interpersonal violent action. In the case of the second individual, there is no doubt that cut marks found on the right humeral shaft are evidence of perimortem trauma. However, I cannot reject the possibility that another situation may have caused these marks on the bone, such as a shark attack (Siegel 1992:213); they could also have occurred during the soft tissue cleaning processes used to separate body parts for use as relics in ancestor worship (Crespo-Torres 2000a, 2002; Rodríguez and Terrazas 2003). At present the skeletal remains in the osteological collections in Puerto Rico, as well as in the rest of the Caribbean, exhibit very low frequencies of skeletal lesions associated with interpersonal violence. These low frequencies are more suggestive of sporadic interpersonal violence than intergroup violence or endemic warfare. The demographic profiles from the four archaeological sites support this conclusion. Artificial cranial deformation practices are reported for skeletal remains from Tibes, Punta Candelero, Paso del Indio, and Maisabel. Most scholars agree that this kind of cultural modification was practiced exclusively by the agro-ceramic groups in the Caribbean area. But, we need to answer at least two important questions: first, which agro-ceramic group or groups introduced this practice into the Caribbean area? Second, what is the real social function or cultural significance of this practice? Some authors (Luna Calderón 1976; Pina 1972; Rivero de la Calle 1949, 1985, 1993) have argued that this practice was introduced by the first agro-ceramic groups from South America that migrated into the Caribbean archipelago, particularly the Cedrosan Saladoids (Rouse 1992:77–85). According to Irving Rouse (1992:80), the Cedrosan Saladoid people migrated into the Antilles as early as the middle of the first millennium b.c. Changes in multiple aspects of the archaeological record, such as differences in ceramic styles, diet, settlement patterns, and physical anthropological remains (e.g., artificial cranial deformation practices and dental morphology), are used by some of Rouse’s followers to support cultural continuity between the Saladoid and the Ostionoid occupations. According to Siegel (1992:363), the site structural data, physical anthropological remains, and 14C evidence from Maisabel support this cultural continuity. Likewise, Budinoff (1991) argues that the dental pathologies and cranial deformation practices present at Maisabel support this “in situ” cultural development. According to Budinoff (1991), some of the skeletons (burials 2, 4, 7, 17, 26, and 30) that showed artificial cranial deformation at Maisabel came from different occupational contexts from the Saladoid (Hacienda Grande) to the Ostionoid. Therefore, for her this cultural continuity is clear. However, the human remains with artificial cranial deformation at Maisabel range from the Cuevas to the Santa Elena styles (Siegel 1992:Table 4.13, 260–263). None of these burials came from the early part

Ancient Bones Tell Stories / 205 of the Cedrosan Saladoid component, represented in Puerto Rico by the Hacienda Grande component (100 b.c.–a.d. 300). On the other hand, Venezuelan archaeologists Mario Sanoja and Iraida Vargas (personal communication 2007) have mentioned that to date no human remains from Saladoid sites in Venezuela, the Saladoid “motherland,” show evidence of artificial cranial deformation. Hence, this cultural modification was not introduced into the Caribbean by the Cedrosan Saladoid people. An alternative hypothesis (Crespo-Torres 2000a, 2005) is that the artificial cranial deformation practice was introduced into the Caribbean by another agroceramic group not related to the Cedrosan Saladoid, the Huecoid (300 b.c.– a.d. 300). In the 1970s Luis Chanlatte Baik and Yvonne Narganes Storde, archaeologists at the Center of Archaeological Research at the University of Puerto Rico, made some discoveries at the La Hueca site that led them to introduce this new early agro-ceramic complex to Caribbean archaeology (Chanlatte Baik 1981, 1995; Chanlatte Baik and Narganes Storde 1980, 1983). The general style of Huecoid ceramic work as well as the elaboration of their lapidary (in particular, the representations of condors) suggests a South American origin for this cultural complex. According to Chanlatte Baik and Narganes Storde, these groups could possibly have origins in the Andean regions of Venezuela or Colombia. Based on these discoveries, Chanlatte Baik and Narganes Storde have argued that contrary to Rouse’s scheme, there were two, almost simultaneous, migrations from South America, which they call Agro-I (or Huecoid) and Agro-II (or Saladoid). In my opinion, the Huecoid people introduced the cranial deformation practice into the Antilles and then it was adopted by the late Cedrosan Saladoid people known as Cuevas (ca. a.d. 400). One characteristic artifact associated with this cultural complex is a biomorph amulet that represents a bird, identified as the Andean condor, holding in its claw a bird and/or a human head. The profile of some of these human heads shows a continuous straight line from the nose tip to the forehead (Figure 9.2b), emphasizing aesthetically the presence of cranial deformation. The absence/presence of artificial cranial deformation to distinguish Saladoid from Huecoid was argued first by Jean-François Durand and Henry Petitjean Roget (1991) in their analysis of a human female skeleton from the site of Morel in the island of Guadeloupe. According to the authors, this skeleton is different from those of the Saladoid for two reasons: first, the skeleton was buried with a funerary necklace and pottery associated with the Huecoid component and, second, the forehead bone shows artificial cranial deformation. Based on these findings, I concur with Chanlatte Baik and Narganes Storde’s point of view that the Huecoid is a new agro-ceramic complex, and I would like to add that the Huecoid people are highly correlated with the introduction of artificial cranial deformation practices into the Caribbean area. Finally, the function of this practice remains unknown, although traditionally

206 / Edwin F. Crespo-Torres it has been conceived as having an aesthetic purpose. At the present, most scholars agree that the practice of artificial cranial deformation was used to identify either those individuals who had power and high social positions in the society (social class or strata distinction) or some form of cultural identity (e.g., ethnicity) among different groups (Hoshower et al. 1995; Munizaga 1992). In my opinion, the late Cedrosan Saladoid people adopted this practice from the Huecoid people to differentiate those individuals with important social status inside the tribe. Then in the shift from tribal to chiefdom society, the artificial cranial deformation was generalized inside the group, since it was used as a marker to differentiate ethnic groups (Crespo-Torres 2000a:212–233).

Conclusion In conclusion, despite the poor state of preservation and conservation presented in the skeletal sample, it was possible to obtain some relevant information regarding bio-cultural aspects of the individuals who were born, lived, and died over eight centuries ago at the Ceremonial Center of Tibes. Unfortunately, the lack of information on the excavations prevented the reconstruction of their social dynamics and mortuary behavior and allowed only partial osteobiographical reconstruction. We hope that in the near future research can be directed at Tibes toward discovering additional burial areas, which would allow not only the reconstruction of mortuary practices but also the application of a bioarchaeological analysis at this important archaeological site. However, despite the limitations mentioned above, the present study provides the following tentative conclusions. First, the demographic profiles from Tibes and Punta Candelero show a similar distribution of age at death, and both present some bias in their sex distribution. On the other hand, Paso del Indio shows a “normal distribution” and at Maisabel, in general terms, adult age predominates over subadult age and sex distribution shows a 1:1 sex ratio in which males lived longer than females. Reexamination of skeletal samples from Maisabel could provide us with the opportunity not only to apply new methods and techniques for the determination of sex and age but also to obtain a more precise demographic profile of this collection. Second, the height profiles from Tibes, Punta Candelero, Paso del Indio, and Maisabel show that stature ranged from 1.58 to 1.60 m (5′2″ to 5′3″) in males and from 1.45 to 1.57 m (4′9″ to 5′1″) in females. Third, caries and heavy dental wear are the dental pathological conditions most frequently found at the four sites mentioned in this study. The consumption of carbohydrate products, such as manioc, maize, and some fruits, is strongly correlated with the high frequency of cariogenic lesions. On the other hand, the high frequency of cases of heavy dental wear is related to multiple factors such as the type of hard food consumed, the presence of gritty or small particles of clay or

Ancient Bones Tell Stories / 207 sand adherence during the food preparation, and, in the case of LSAMAT, the use of teeth as a tool. Fourth, cases of shovel-shaped incisors were recorded in the populations from Tibes, Punta Candelero, and Paso del Indio. Unfortunately, the study of human remains from Maisabel does not make reference to either the presence or absence of this dental morphological trait. Therefore, future dental analysis of the collection from Maisabel is needed to assess this trait. Fifth, the bone pathologies found most frequently in the four skeletal collections are periostitis and osteoarthritis. However, the first of these is more frequent at Paso del Indio, affecting basically the subadult group, and is probably related to a nutritional stress situation. The second is more frequent in adult groups and is associated with bone degenerative processes related to old age. Sixth, the demographic profile and the low evidence of trauma related to interpersonal violence are more suggestive of sporadic interpersonal violence than endemic warfare. Seventh, artificial cranial deformation, as a cultural modification, is present at Tibes, as well as at Punta Candelero, Paso del Indio, and Maisabel. Introduced into the Caribbean by the Huecoid people, this cultural modification is adopted by late Saladoid groups with a tribal social organization to differentiate high-ranking individuals inside the group. Later, in the transformation process from tribal to chiefdom society the practice of artificial cranial deformation is generalized into the group as a marker of ethnicity. Finally, the demographic profile, pathological conditions, and cultural modifications present at Tibes, Punta Candelero, Paso del Indio, and Maisabel are also documented in other Caribbean islands, such as Cuba (Rivero de la Calle 1949, 1985, 1987; Vento and González 1996), the Dominican Republic (Luna Calderón 1976, 1988, 1993, 2001; Morbán et al. 1976), Jamaica (Santos et al. 2002), and the Lesser Antilles (Baetsen 1999; Hoogland 1996; Righter 2002; Versteeg and Rostain 1996; Versteeg and Schinkel 1992). However, it is necessary to develop more comparative studies from a bioarchaeological perspective (Buikstra 2006) in the areas of skeletal biology and paleopathology in order to determine diachronic and/or synchronic regional differentiation among the prehistoric populations of the Caribbean. The bones and teeth of the early inhabitants of Tibes have spoken to us about part of their lives, but they have more stories to tell us; stories that still lie buried in the soil of this important archaeological site.

Acknowledgments I would like to express my deep gratitude to Dr. Antonio Curet for the invitation to contribute to this volume, and also for his and Lisa Stringer’s assistance in editing and commenting on the manuscript. I also would like to make a special refer-

208 / Edwin F. Crespo-Torres ence to the late Rafael “Churumba” Cordero Santiago, Mayor of the Autonomous Municipality of Ponce, for his particular interest in and support of the research of our ancient history. I also thank the members of the Ceremonial Center Advisory Board, in particular Dr. Ray Petty, for his dynamism and commitment to the culture and history of our Boriquén. And, finally, I acknowledge all the personnel of the Ceremonial Center of Tibes, particularly the late José Lugo Santiago, Irma L. Zayas Alvarado, and Carmen Martínez Ajá.

10 Bone Chemistry and Paleodiet at the Ceremonial Center of Tibes William J. Pestle

The reconstruction of past human diet is one of the most basic concerns of the archaeologist. Food production and distribution, besides being essential to survival, are defined by, and help to define, a broad range of social, cultural, and political processes. In almost every society, food forms a key part of social and economic relations at multiple levels, as well as serving as a symbol related to different aspects of culture (e.g., ethnicity, political power, religion) and as an expression of forms of social identity (for a review on this topic see, for example, Ashley et al. 2004; Farb and Armelagos 1983; Harris 1985; Mintz and Du Bois 2002). In particular, the use and control of food can become an arena where power and prestige are negotiated (Appadurai 1981; Danforth 1999; Hastorf and Johannessen 1993; Ross 1987). Thus, it is important to recognize that food is a dynamic entity in a wide range of social relations in both private and public spheres. Within the cultural historic framework of Caribbean archaeology, however, most studies have treated diet, subsistence, and food only as fossil indexes of chronological or cultural periods. Rainey (1935, 1940) was the first to do so when he used diet as a marker for shifts in ethnic identity in Puerto Rico. During his excavation in Ponce, Rainey noted that Cedrosan Saladoid subseries assemblages contained high concentrations of crab claws, while those from the later Elenan and Chican Ostionoid subseries contained high quantities of shellfish instead. This led him to define these deposits as belonging to successive peoples of the “Crab” and “Shell” cultures, respectively. In more recent years, deFrance (1989) conducted a more detailed analysis of Cedrosan Saladoid and Elenan Ostionoid assemblages from Maisabel in order to investigate and test Rainey’s dichotomy of Crab and Shell cultures. In the 1980s and 1990s, Yvonne Narganes Storde (1985,

210 / William J. Pestle 1993a, 1993b) also began to conduct more refined zooarchaeological studies intended to identify differences in the diet of distinct cultural groups (Huecoid and Cedrosan) in the early Ceramic age. In all of these cases as well, food remains were used as a diagnostic to define the different assemblages of different, but in this case contemporary, “peoples.” While valuable, these studies have failed to view food as an active element in the structuring of interpersonal relationships of the society in which it was being consumed. This reduces food to little more than a passive marker of people, rather than a potentially useful tool for the reconstruction of past social processes. In order to distill meaningful and accurate insights about the broader social dimensions of food in any given past society, one must first employ an analytical methodology that is capable of reconstructing past subsistence patterns with sufficient fidelity and at a satisfactory resolution. Traditionally, paleodietary reconstruction has involved either a reliance on possibly inaccurate or slanted historical sources or the indirect archaeological study of diet through the analysis of food remains (zooarchaeology, paleoethnobotany) or nutrition-related pathologies in human skeletal remains (paleopathology). While these techniques are all capable of providing valuable dietary insights, all also possess theoretical and/or methodological shortcomings that can result in overly coarsely grained and/or misleading reconstructions of past human diet. The study of food remains (either faunal or floral) is, perhaps, the most common means of paleodietary analysis. When carried out diligently, such techniques can provide detailed information on the different types of food resources exploited, quite often with species-specific resolution. This resolution is extremely useful since habitats and possible methods of exploitation can be inferred from the identification of particular species, and when combined with other types of information, such analyses can provide insights into matters such as the control of labor. However, the use of techniques involving the indirect study of diet through food remains can produce nonrandomly biased or misleading results because of the differential survival and preservation of different classes of food (Pearsall 2000). In the Caribbean, as in the neotropics in general, where many important foodstuffs (e.g., manioc) are composed mainly of soft plant tissues that survive poorly or not at all, this is of particular concern (Newsom 1993). As floral and faunal evidence is rarely recovered in the relative quantities in which it was consumed, the conclusions derived from such evidence are necessarily qualitative rather than quantitative (Ambrose 1993:59). Moreover, it is important to remember that the scale of the data derived from the study of food remains is at the level of a community or group rather than that of an individual, that the temporal resolution provided by such evidence can be quite variable, and that the data it can produce can be misleading due to many cultural and natural formation processes that may have

Bone Chemistry and Paleodiet at the Ceremonial Center of Tibes / 211 affected its evidentiary basis (Barberena and Borrero 2005). Thus, indirect means of dietary reconstruction will tend to fail in the detection of subtle differences in diet between individuals. Put differently, the study of the actual food remains provides us with the menu rather than the meal (Ambrose 1993:82, 85). The study of health indexes in human bones is another means of acquiring information about past peoples’ diets. Since many aspects of health are related to nutrition, skeletal health indicators and pathologies can provide indirect information about the types and quantities of foods consumed. Prevalence rates for skeletal pathologies such as cribra orbitalia, porotic hyperostosis, and linear enamel hypoplasias; the overall disease load in a population; and broader indicators such as stature have all been argued to serve as effective proxies for nutritional well being (Buikstra 1992; Goodman et al. 1988; Roberts and Manchester 1999). While absolute statements about nutrition are not possible as a result of this type of analysis, relative statements based on changes through time or differences between individuals resulting from differences in status are possible. The inferences derived from such studies are necessarily broad, however, and can be fraught with potential difficulties given the multiple etiologies of the pathologies under discussion and the fact that the individuals under study may not be representative of the population at large (Ambrose 1993; Wood et al. 1992). Furthermore, such analysis can really only be used to infer nutritional inadequacy rather than the presence or relative contribution of any actual dietary items (Schwarcz and Schoeninger 1991:284). In light of the methodological shortcomings of the techniques most often employed in the study of diet in the prehistoric Caribbean, this chapter presents the results of a pilot study undertaken to assess the feasibility of instead using biogeochemical analytical techniques to reconstruct the subsistence patterns of the prehistoric inhabitants of Puerto Rico. The analysis of the stable isotopes of two light elements (carbon and nitrogen) in both the organic and mineral phases of human bone was employed to facilitate the direct dietary reconstruction of four individuals interred within the Ceremonial Center of Tibes. This work forms part of a larger ongoing project employing stable isotope analysis of Puerto Rican skeletal materials of varying ages and depositional histories that aims to define the relationship between subsistence and the in situ development of social complexity in prehistoric Puerto Rico. It is hoped that the use of this stable isotope technique, when combined with multisource mixture modeling, will produce data of sufficient resolution and quality to allow for the reconstruction of the possible dietary dimensions of fine-grained social processes. If successful, it is hoped that the application of this technique will allow for the detection of subtle, individual-level changes in diet possibly reflective of a diachronic shift in sociocultural organization.

212 / William J. Pestle

Dietary Reconstruction Using Stable Isotopes At its simplest, stable isotope analysis of paleodiet is predicated on the notion that you are, isotopically, what you eat. Stable isotope analysis of prehistoric human diets involves the comparison of carbon and nitrogen isotope ratios from human consumer tissues (in this case the organic and inorganic phases of bone) with those of the edible portions of possible food sources. While the relationship between the isotopic values of consumer and source tissues is far from one-to-one due to fractionation factors, the results of this analysis can allow one to determine, with a healthy degree of imprecision, the relative importance that various foodstuffs may have had in a past individual’s diet. Stable isotopes are nonradioactive variants of an element differing from the most abundant isotope in the number of neutrons in the nucleus (Fry 2006). Due to the differences in their atomic masses, different isotopes of the same element can behave differently both kinetically and thermodynamically during chemical reactions, thereby resulting in a characteristic patterning of these isotopes in organisms and their tissues (Fry 2006; Schwarcz and Schoeninger 1991; Urey 1947). For the purposes of dietary reconstruction, the stable isotope ratios of two elements, carbon and nitrogen, are of principal interest, because these elements are taken up by plants from soil, water, and air and are passed on in a known fashion through various consumers in a food web (Ambrose 1987; Bender et al. 1981; Delwiche et al. 1979; DeNiro 1987; DeNiro and Epstein 1978, 1981; Schoeninger 1985; Schoeninger and DeNiro 1984; van der Merwe 1982; Vogel and van der Merwe 1977). The reconstruction of prehistoric subsistence through stable isotope analysis is today a well-established technique, with some four decades of supporting literature. The technique has been employed in ecological studies since the 1960s (Parker 1964), and the possibility of its usefulness for the study of past human diets was realized soon after its inception (Hall 1967). While techniques have been refined and improved since their earliest practical applications (Bender et al. 1981; Chisholm et al. 1983; Chisholm et al. 1982; Delwiche et al. 1979; DeNiro 1985; DeNiro and Epstein 1978, 1981; Klepinger 1984; Schoeninger 1985; Schoeninger and DeNiro 1984; Schoeninger et al. 1983; Sullivan and Krueger 1981, 1983; Tieszen et al. 1983; van der Merwe 1982; van der Merwe et al. 1981; Vogel and van der Merwe 1977), the theoretical basis for the analysis remains the same: predictable variation in the isotopic signatures of certain classes of plant and animal foodstuffs (resulting from their preferred sources of food [and thus carbon and nitrogen] and their position in their respective food webs) can make possible the reconstruction of the relative quantitative contribution each class of foodstuff made to the diet of an analyzed consumer. More recent studies (e.g., Ambrose and Norr 1993) have refined the technique, as well as our understanding of the underlying

Bone Chemistry and Paleodiet at the Ceremonial Center of Tibes / 213 processes thereof, to the extent that now both protein and energy (carbohydrate, lipid) portions of an individual’s diet can be reconstructed more or less independently. In stable isotope studies, isolated and purified samples (preparation methods are discussed below) are converted to a gaseous form and analyzed using an isotope ratio mass spectrometer (Ambrose 1993:67–71; Brenna et al. 1997). The results of this analysis are expressed as the ratio of the heavier to the lighter isotope (13C to 12C in the case of carbon, 15N to 14N in the case of nitrogen) and are typically presented using a delta notation (d) in parts per thousand (per mil or ‰) relative to an international environmental standard. Sample delta values for carbon and nitrogen are determined using the following formulas: d 13C =

⎡( ⎣

13

⎤ × 1000

C/12C)sample − (13C/12C)PDB (13C/12C)PDB



and d 15N =

⎡( ⎣

15

⎤ × 1000

N/14N)sample − (15N/14N)AIR 15

14

( N/ N)AIR



The standard employed for carbon is the Vienna Pee Dee marine fossil formation (PDB, Craig 1957), which contains far more 13C than most human foodstuffs or tissues, thus rendering most archaeological d 13C values as negative numbers. The nitrogen standard is atmospheric nitrogen (AIR, Mariotti 1983), which is more depleted in 15N than most archaeological samples, with the result that most reported values in archaeological studies are positive numbers. Broadly, d 13C values in flora discriminate exclusively between plants of the so-called C3 (trees, fruits, tubers, temperate grasses) and C4 (maize, sugarcane, sorghum, amaranths) photosynthetic pathways, with so-called CAM (Crassulacean Acid Metabolism, e.g., cacti or pineapple) and marine plants occupying the middle ground between the C3–C4 isotopic extremes (Sage et al. 1999; Smith and Epstein 1971). The cause of this isotopic difference can be found in the varying degree to which the different plant classes fix 13C available from atmospheric sources of CO2 (d 13C = −7‰). As a result of these isotopic/photosynthetic differences, C3 plants have d 13C values averaging −26‰, whereas C4 plants average −12‰ (Bender et al. 1981; Smith and Epstein 1971). The tissues of the consumers of these plants, be they animals or humans, continue to reflect the isotopic signatures of their plant-based diet onward through the food chain, with, for example, the tissues of exclusive consumers of C4 plants having a much more enriched (less negative) d 13C value than that of an exclusive C3 consumer. Values of d 13C can also be used as a marker of the consumption of marine

214 / William J. Pestle foods, as the bicarbonate ingested by most marine species is enriched by roughly 7‰ over atmospheric carbon dioxide, with the exception of reef and seagrass dwellers, which are enriched even more (4–7‰ greater still) than their pelagic marine neighbors (Chisholm et al. 1982; DeNiro and Epstein 1978; Fry et al. 1982; Schoeninger and DeNiro 1984; Schoeninger et al. 1983). Given the proximity of their d 13C values, diets that incorporate both marine foods and C4 terrestrial foods can result in problematic interpretive issues. Signatures of d 15N, on the other hand, can differentiate between legumes (averaging a d 15N of 1‰) and non-nitrogen fixing plants (an average of 3‰) (Delwiche et al. 1979; Schoeninger and DeNiro 1984). Nitrogen isotope signatures also pattern strongly with trophic level, with a stepwise increase of approximately 3–5‰ between the tissues of consumers at successive trophic levels (herbivores, primary carnivores, secondary carnivores) (Schoeninger 1985; Schoeninger and DeNiro 1984). This latter effect is of particular use for archaeological studies of past diet as it allows for the discrimination of marine vs. terrestrial diets. This determination is possible as a result of the relatively “longer” trophic chains seen in marine ecosystems, which result in the fact that marine species tend to exhibit more enriched d 15N signatures than do their terrestrial equivalents (Bösl et al. 2006:297). Most of the early archaeological applications of stable isotope analysis focused solely on the isotopic makeup of collagen, the organic (proteinaceous) fraction of human bone. This focus was a result of the persistence of the linear mixing model, which posited that consumed materials, and thus consumed isotopes, were incorporated in equal amounts in any and all tissues of the consumer (van der Merwe 1982; White and Schwarcz 1989). Since, under this model, all tissues were believed to reflect faithfully the isotopic makeup of one’s diet, collagen from archaeological samples, because it presented fewer problems in terms of diagenetic alteration, became the preferred material for analysis. However, more recent research, including controlled diet experiments, has made clear that a different model, which combines portions of linear mixing with a macronutrient routing model (Chisholm et al. 1982; Krueger and Sullivan 1984), better reflects the way in which the constituent portions of foodstuffs are incorporated into a consumer’s tissues. Ambrose and Norr’s landmark 1993 study on this matter verified that different components of a consumer’s diet were preferentially incorporated into and reflected by different tissues of that consumer. Thus, collagen in human bone primarily reflects the protein portion of that individual’s diet (not the whole diet as had been previously assumed), whereas the apatite (mineral phase) of the same human bone reflects the entire diet, including both dietary protein and the dietary energy component (carbohydrates and lipids). Additionally, Ambrose and Norr (1993) found that the spacing between d 13C values in collagen and apatite, D13Cap−co, can be of great utility in understanding the amount of protein in the diet as well as the relative amounts and isotopic signature of energy and protein foodstuffs.

Bone Chemistry and Paleodiet at the Ceremonial Center of Tibes / 215 A final key result of Ambrose and Norr’s (1993) study was their demonstration of the inappropriateness of applying a supposed universal fractionation factor to d 13C values (normally +5‰) of consumer collagen in an attempt to determine the carbon isotope value for that consumer’s proteinaceous food sources. While, on the one hand, D13Cap−diet is effectively constant at or around +9.4‰, based on the isotopic differences or similarities within and between the energy and protein portions of a consumer’s diet (e.g., C4-derived protein and C3-based energy), the degree of fractionation between consumer collagen and dietary protein (or D13Cco−protein) can range from +12.6‰ to −0.5‰ (Ambrose and Norr 1993). Thus, it is not possible to simply subtract 5‰ from consumer collagen d 13C to arrive at the d 13C value of the foods they ate. Nonetheless, with a detailed knowledge of the isotopic character of potential food sources, and in light of controlled diet experiments, it is still possible to reconstruct both the protein (meat) and energy (carbohydrate) components of a consumer’s diet independently and with notable accuracy as long as caution is exercised in the application of correction/fractionation factors.

Previous Caribbean Isotopic Studies To date, only six isotopic studies have been undertaken of prehistoric subsistence in the Caribbean (Keegan 1985; Keegan and DeNiro 1988; Norr 2002; Schoeninger et al. 1983; Stokes 1998; van Klinken 1991). The four earliest studies were conducted when the linear mixing model (see above) was still en vogue and thus relied on the analysis of bone collagen alone for the reconstruction of the whole diet. This methodological shortcoming serves to substantially, if not totally, mitigate the usefulness of their results. Moreover, van Klinken (1991) included reported results for samples whose collagen C:N ratios (discussed below) were suggestive of a high degree of postmortem diagenetic alteration. Fortunately, Stokes (1998) included in her dissertation a reanalysis of many of the same individuals previously analyzed by Keegan, DeNiro, and van Klinken, in addition to a large number of individuals from other sites throughout the Caribbean. Stokes’s work thereby greatly augments and normalizes the comparative data set for the present study. Norr (2002) adds to this comparative corpus the results of an analysis of more than 20 individuals from the Tutu site in the U.S. Virgin Islands.

Methods and Materials In August 2004, with funding from the H. John Heinz III Foundation, samples of human bone and teeth were extracted from skeletal material from four prehistoric Puerto Rican sites: Maruca, Punta Candelero, Paso del Indio, and Tibes (Table 10.1). Each site is located in a geologically and ecologically distinct portion of the island and, in general, the materials from them belong to different

216 / William J. Pestle Table 10.1. Individuals sampled for stable isotope analysis Site

Date

Individual

Age

Sex

Paso del Indio Paso del Indio Paso del Indio Paso del Indio Tibes Tibes Tibes Tibes Punta Candelero Punta Candelero Punta Candelero Punta Candelero Maruca Maruca Maruca Maruca

a.d. 900–1200 a.d. 900–1200 a.d. 900–1200 a.d. 900–1200 a.d. 400–1200 a.d. 400–1200 a.d. 400–1200 a.d. 400–1200 a.d. 400–600 a.d. 400–600 a.d. 400–600 a.d. 400–600 ca. 3000 b.c. ca. 3000 b.c. ca. 3000 b.c. ca. 3000 b.c.

PDI 6.0002 PDI 6.0050 PDI 6.0056 PDI 6.0074 E-48 E-13 ES-3 CE-5 PC-33 PC-36 PC-46 PC-56 2B 4 9 10

Adult Adult Adult Adult 35–40 35–40 Adult 35–40 Adult Adult Adult Adult ? ? ? ?

Female Female Male Male Female Male Male Female Male Male Male Male ? ? ? ?

cultural/chronological periods. Individuals were included in this study only when both age and sex could be assessed reliably, and the skeletal elements selected for sampling were required to be robust and dense. Any evidence of heat alteration, discoloration, preservative agents, consolidants, or fungal activity disqualified individuals or elements from the present study. Gloves and masks were worn throughout the sample extraction process, and sampling instruments were cleaned between individuals with dilute bleach in order to decrease the potential for crosscontamination of samples. While sampling requirements have since been revised, at the time of this pilot project two teeth and approximately 5 g of bone per individual were extracted. Samples were later subdivided for three types of analysis (stable isotope, trace element, and mtDNA) in a laboratory of the Anthropology Department of The Field Museum of Natural History, Chicago.1 Funding for the pilot study limited sample size to four individuals per site (for a total of 16 individuals), rendering the resultant data too limited for anything but the most tentative conclusions, but useful for assessing the feasibility of further study. In the context of the present work, however, it was decided to go marginally further and use the present, admittedly limited, data set to formulate some working hypotheses about trends in Puerto Rican paleodiet by which we might direct our future research on the topic. Human bone samples from the 16 individuals (including the four individuals from Tibes) were analyzed at the Environmental Isotope Paleobiogeochemistry

Bone Chemistry and Paleodiet at the Ceremonial Center of Tibes / 217 Laboratory of the Department of Anthropology at the University of Illinois Urbana-Champaign in 2004 and 2005 using the sample preparation and analytical methods described in Ambrose et al. (2003). Broadly, this process involves the mechanical cleaning and grinding of samples, followed by a division of the samples into fractions used for separate analysis of the mineral and collagenous portions of the bone. These split samples are subsequently isolated and purified (a process that involves demineralization, dehumification, filtration, and gelatinization for collagen and organic and labile carbonate removal for apatite) prior to their separate combustion and analysis via mass spectrometry. As the principal goal of the pilot study was to assess the usefulness of stable isotope analysis for studying paleodiet in prehistoric Puerto Rico, assurance was required that samples under analysis were sufficiently unaltered following their prolonged period of burial. As the chemical interactions of bone with other body tissues and burial soils (diagenesis) can alter bone isotopic values and thus produce misleading dietary results, it is necessary in any instance of stable isotope analysis to assess the character and quality of the preserved osseous materials. To this end, percentage yield collagen, percentage yields carbon and nitrogen, and carbon to nitrogen ratio (C:N), the most commonly employed sample fitness standards, were calculated for all samples in order to assess their suitability for further study (Ambrose 1990; DeNiro 1985; van Klinken 1999). The use of a continuous flow isotope ratio mass spectrometer allowed for the detection of both d 13C and d 15N values for bone collagen and d 13C values for bone apatite from all samples in question. The results of this step of the analysis provide one with the isotopic signature of a consumer, but in order to arrive at the isotopic value of that individual’s diet, one must employ correction factors to compensate for the isotopic changes brought about by chemical fractionation between consumer and consumed. Fractionation appears as an enrichment or depletion of the isotopic signature of a consumer’s tissue relative to the isotopic value of the foods being consumed, with different tissues and different organisms fractionating their dietary isotopes differently. As discussed above, in the case of collagen d 15N and apatite d 13C, the process of correcting for fractionation is a fairly simple one. To arrive at the d 15N of an individual’s diet, one can simply subtract 3‰ from the d 15N value derived from the collagen in order to account for the trophic level increase between the human under analysis and his/her diet, although published estimates do range between 1.6‰ and 5‰ (Ambrose 2000; Ambrose et al. 2003; Bocherens et al. 1995; Fry 1991; Hedges and Reynard 2007; Schoeller 1999; Schoeninger 1985; Schoeninger and DeNiro 1984; Schwarcz 1991; Schwarcz and Schoeninger 1991; Tuross et al. 1994). Similarly, in the case of apatite d 13C, a −9.4‰ correction factor to the consumer d 13C signature allows one to arrive at the d 13C value of the consumer’s whole diet (Ambrose and Norr 1993). However, in the case of bone collagen d 13C,

218 / William J. Pestle the matter of correcting for fractionation is substantially more complex. This complexity is the result of a compound relationship between the isotopic value of dietary protein, energy, and the relative amounts of each in the consumer’s diet, with all of these playing a role in determining the resulting spacing between analyzed collagen d 13C and the d 13C value of dietary protein. For the purpose of the present analysis, we have employed a very conservative collagen to protein spacing value of +1.1‰. The use of this spacing value was determined to be appropriate for the present study based on the observed strong negative correlation (−1.0) in the published data of Ambrose and Norr (1993) between D13Cap−co and the isotopic spacing from the d 13C of collagen to that of dietary consumables. We used this observed correlation to generate a predictive formula for deriving an appropriate collagen–diet spacing value based on the D13Cap−co of a given sample. Use of this formula resulted in the +1.1‰ value employed herein. The use of a +1.1‰ spacing will, if anything, tend to underestimate the contribution of traditionally underappreciated foodstuffs (freshwater and terrestrial fauna) to the Tibes individuals’ diets. Even so, as is discussed below, the results of this analysis suggest that such sources actually dominated the protein portion of these diets, and thus the +1.1‰ value employed would appear to represent an extremely conservative estimate of d 13C collagen to protein spacing. In order to establish a set of baseline values for the dietary options to which the prehistoric human consumers of Tibes may have had access, it was also necessary to compile isotopic data for a range of potential foodstuffs. As the pilot study did not include funds to analyze any nonhuman (floral and faunal) materials, isotope values for possible foodstuffs were compiled from previously published results (Keegan 1985; Keegan and DeNiro 1988; March and Pringle 2003; Newsome et al. 2004; Stokes 1998). Due to discrepancies in the analytical methodology employed by these various studies, the lack of published data for many potential foodstuffs from the immediate area of Tibes, and the inclusion in the present study of isotopic data from species and individuals that may not faithfully represent the values of similar foods consumed at Tibes, the data compiled in this exemplar food web are best viewed as tentative and preliminary. A correction factor of +1.5‰ d 13C was applied to any modern food samples used in the construction of the ancient food web in order to account for the enrichment of atmospheric 13C brought about by the recent burning of fossil fuels (Keeling et al. 1979; Marino and McElroy 1991). As well, corrections for the isotopic differences between consumable (flesh) and tested (bone or shell) tissues (for fish, −3.7‰ d 13C and +1.7‰ d 15N; for all other animals, −2.3‰ d 13C and +0.6‰ d 15N) have been applied as appropriate (DeNiro and Epstein 1978, 1981; Keegan and DeNiro 1988; van der Merwe 1982). To further enhance the usefulness of stable isotope analysis for the reconstruc-

Bone Chemistry and Paleodiet at the Ceremonial Center of Tibes / 219 tion of the protein portion of the diet of the Tibes individuals, we also made use of a sophisticated multisource mixing analysis software called IsoSource (Phillips and Gregg 2003; Phillips et al. 2005). This software has only recently begun to be used in the reconstruction of past dietary patterns (Bocherens et al. 2005; Drucker and Henry-Gambier 2005; Newsome et al. 2004). Originally developed by researchers at the U.S. Environmental Protection Agency for ecological and geological applications, this software resolves the complicated computational issues that lie behind analysis of mixing equations where the number of potential sources is greater than n + 1, where n equals the number of informative isotope ratios (e.g., d 13C) possessed for both a consumer and all its potential food sources (Phillips and Gregg 2003; Phillips et al. 2005). The application of this technique, while failing to provide a silver bullet or definitive dietary combination for a past consumer, determines the range of (mathematically) feasible contributions of each potential food source to a given individual’s diet. This is accomplished through a process of simulation wherein all possible combinations of proposed foodstuffs, and their accompanying isotopic signatures, are tested against the observed isotopic values of the consumer under analysis: “In this method, all possible combinations of each source contribution (0–100%) are examined in small increments (e.g., 1%). Combinations that sum to the observed mixture of isotopic signatures within a small tolerance (e.g., ±0.1‰) are considered to be feasible solutions, from which the frequency and range of potential source contributions can be determined” (Phillips and Gregg 2003:261). Thus, while without this software the two isotope systems used in dietary reconstruction (carbon and nitrogen) might allow for the determination of the relative contributions of three (n + 1, where n = 2) different foodstuffs, the use of multisource mixture modeling facilitates the determination of the dietary contributions of many more food sources, a situation that better mirrors the dietary variety available in the protein portion of the prehistoric Puerto Rican food web. The percentage of C4/CAM plants in the diets of the individuals under study was calculated using the following formula (Ambrose et al. 1997; Schwarcz and Schoeninger 1991): %C4 = ((−25−(d 13Cap−9.4‰))/15) × 100 Due to uncertainty introduced by individual variation, instrumentation, and temporal factors, as well as the possibly confounding effect brought about by the presence in the diet of substantial quantities of marine foodstuffs, the accuracy of the results of this calculation of dietary percentage C4/CAM should be considered to be ±10 percent (Ambrose et al. 2003). Finally, in order to assess the chronological dimensions of dietary change (or

220 / William J. Pestle Table 10.2. Quality standards and assessment of individuals interred at Tibes Individual Standarda E-48 E-13 ES-3 CE-5

% Yield Collagen

% Yield Carbon

% Yield Nitrogen

Carbon:Nitrogen Ratio

>.5, >3 0.57 2.32 0.48 2.29

>4.5, >13 29.71 43.53 17.13 39.49

>.5, >4.8 9.93 14.87 5.50 13.52

2.8–3.6 3.49 3.42 3.63 3.41

a

The first value is the minimum accepted for continuation of analysis; the second value is the preferred amount.

the lack thereof ) at Tibes, collagen samples from all of the individuals discussed herein were AMS dated at the Accelerator Mass Spectrometry Laboratory at the University of Arizona, Tucson.

Results Of the samples extracted from the four individuals buried at Tibes, three (E-48, E-13, and CE-5) contained a satisfactory amount of collagen (averaging 1.7 wt%) and possessed C:N (carbon to nitrogen) ratios (3.44 average) at or exceeding acceptable standards (Table 10.2). In addition, the fourth individual (ES-3) presented a collagen yield of 0.48 wt% and a C:N ratio of 3.63, both of which deviate only minimally from desired/accepted levels. Due to the small degree of the departure of ES-3 values from accepted standards, we believed the inclusion of this individual in the present preliminary study to be warranted. These results demonstrate that the Tibes samples were sufficiently unaltered by diagenetic processes and that their resulting isotopic signatures can be considered to be the result of bona fide dietary intake rather than taphonomic alteration. Overall, 75 percent of the samples from the four sites under study (Maruca, Paso del Indio, Punta Candelero, and Tibes) met or exceeded standards for sample fitness, giving us great hope for the future expansion of this study. The isotopic composition of bone collagen and apatite from the Tibes individuals and the predicted isotopic values for the protein and energy portions of their diets are provided in Table 10.3, as are the results of AMS dating of bone collagen from these individuals. The d 15Nco values for these consumers were moderately enriched, ranging from 9.07‰ to 10.37‰, while the d 13Cco values extended from −17.15‰ to −17.72‰. Both of these isotopic indicators would appear to indicate the possibility of an unanticipated source or sources of dietary protein for the individuals interred at Tibes. Moreover, the d 13Cap values for these individuals (which range from −8.01‰ to −9.36‰) are notably enriched, and the D13Cap−co

d Nco (‰)

9.8 10.37

10.32 9.07 9.89

Individual

E-48 E-13

ES-3 CE-5 Mean

15

7.32 6.07 6.89

6.8 7.37

d 15N protein (‰)

–17.15 –17.72 –17.36

–17.37 –17.21

d Cco (‰)

13

Δ Cap–co (‰) 8.26 8.16

7.79 9.71 8.48

≤–18.47 ≤–18.31

≤–18.25 ≤–18.82 ≤–18.46

13

d 13C protein (‰)

–9.36 –8.01 –8.88

–9.11 –9.05

d Cap (‰)

13

–18.76 –17.41 –18.28

–18.51 –18.45

d 13C diet (‰)

41.6 50.6 44.78

43.27 43.67

% C4/ CAM

1392 ± 42 1428 ± 42 n/a

1366 ± 44 1352 ± 43

C Age, Years b.p.

14

(cal a.d. 623: cal a.d. 686) 1 (cal a.d. 641: cal a.d. 691) 0.893963 (cal a.d. 750: cal a.d. 762) 0.106037 (cal a.d. 615: cal a.d. 664) 1 (cal a.d. 599: cal a.d. 653) 1 n/a

Calibrated Dates (1s age range) Area under Distribution

Table 10.3. Analyzed stable isotope ratios of individuals interred at Tibes, predicted isotopic values of diets, and AMS dates

222 / William J. Pestle Table 10.4. Food web groupings employed for paleodietary reconstruction Group Gastropods Sea turtles Marine fish Bivalves Marine crabs Birds Freshwater fauna Pinnipeds Terrestrial fauna C3 plants C4/CAM plants

Mean d 13C (‰)

SD d 13C (‰)

Mean d 15N (‰)

SD d 15N (‰)

–9.89 –9.95 –10.13 –12.81 –15.19 –18.50 –18.88 –19.20 –23.76 –24.86 –9.52

3.33 5.63 2.93 7.17 5.17 3.88 2.14 0.80 1.76 .60 .76

2.42 5.85 7.57 1.43 5.03 7.19 7.19 19.10 6.61 4.07 1.65

1.89 1.88 1.59 0.91 1.37 3.40 1.49 1.20 1.69 1.67 .66

spacing values (all greater than 7.7‰) are fairly high, perhaps as the result of a large amount of C4/CAM-derived dietary carbohydrates, a possibility detailed below. The predicted isotopic signatures for the diets of the Tibes individuals can be compared to the isotopic values of possible dietary elements using the individual carbon and nitrogen isotope values of a wide range (n = 83) of potential foodstuffs, as gleaned from published sources. The values provided in Table 10.4 reflect the isotopic values of these potential foodstuffs with appropriate correction factors for their respective dates and tissue type already having been applied. Given the enormous number of potential individual foods, for the purpose of dietary reconstruction, the individual species or genera of foodstuffs have been grouped by bio-ecological niche in order to produce the 11 categories of consumables represented in Table 10.4.

Discussion Focusing first on the results of the analysis of bone collagen, both the carbon and nitrogen results require discussion. The moderately enriched d 15N signatures (averaging 9.89‰) observed in the collagen of the Tibes human samples speak to the high trophic level that these individuals must have occupied within the region’s prehistoric food web. In fact, the d 15N values of the human samples were on par with the upper range of values observed in typically high trophic level species of marine and freshwater fish and were outstripped only by those of the marine seals (Figure 10.1). This is not to say that the Tibes humans were as carnivorous as

Bone Chemistry and Paleodiet at the Ceremonial Center of Tibes / 223

Figure 10.1. Values of d 15N from Tibes humans and possible elements of protein food web.

barracuda, sharks, or seals, as a little meat goes a long way isotopically, but rather that there was a substantial portion of (high trophic level) animal protein in the Tibes humans’ diets (Ambrose et al. 2003; Ambrose and Norr 1993). The potential source or sources for this high trophic level protein is clarified by the results of the d 13Cco analysis, which produced surprisingly depleted (negative) carbon isotope ratios, averaging −17.36‰. From an examination of the proteinaceous foodstuffs potentially available to the Tibes consumers, it is evident that their d 13C values were likely pulled toward more negative values by the inclusion in the diet of some substantial amount of high trophic level (high 15N) and 13C-depleted protein. Based on the reconstruction of the local food web provided in Table 10.4, the list of likely candidates for this protein source includes freshwater fauna (mean d 15N 7.19, mean d 13C −18.88), birds (mean d 15N 7.19, mean d 13C −18.5), terrestrial fauna (mean d 15N 6.61, mean d 13C −23.76), and pinnipeds (mean d 15N 19.1, mean d 13C −19.2). By utilizing only traditional means of isotopic dietary reconstruction, there remains a notable lack of precision in determining the relative amount of protein that may have been derived from each of these potential sources. All that can be said is that any or all of these sources played an important role in the provision of dietary protein. To some degree, this imprecision is the result of the breadth of the food web

224 / William J. Pestle groupings employed in this analysis and the resulting overlap of the contrived bioecological food groups (Table 10.4). The variation in isotope values within each grouping is likely the result of a number of factors (inappropriate grouping, heterogeneity in geography, environment, or diet for specimens analyzed) and is likely further compounded by small sample size. In any instance of isotopic analysis of diet wherein the various potential food sources show substantial overlap, some degree of ambiguity will necessarily result. However, another source of uncertainty in the reconstruction of the protein portion of the Tibes individuals’ diets, that which results from the sheer multitude of possible food sources, can be minimized by employing multisource mixture modeling software such as IsoSource. As discussed above, this method allows for the resolution of individual source contributions when the number of potential sources is greater than n + 1, where n equals the number of informative isotope ratios (Phillips and Gregg 2003; Phillips et al. 2005). While such techniques do not provide absolute solutions for the relative makeup of an individual’s diet, they do allow determinations of major and minor contributors to be made. The bounded results of the IsoSource analysis of the Tibes individuals’ diets are presented in Table 10.5 (as above, this analysis was undertaken using the very conservative collagen to protein spacing value of +1.1‰). For these purposes, the 1st percentile values are taken to represent a reasonable minimum amount of dietary protein derived from each individual source, the 99th percentile values represent a maximum potential contribution from each source, and the mean values represent the most likely amount of protein derived from each source in that individual’s diet. The mean values of the relative contributions of each of the potential sources (summing to 100 percent) are also represented graphically in Figure 10.2. All told, the results of this mixture modeling paint a rather surprising picture of prehistoric Puerto Rican protein consumption. First, the results suggest a large degree of dietary breadth for the Tibes individuals, in that many ecological niches (rivers, land, ocean, even the air, in the form of birds) were exploited as fruitful sources of dietary protein. Second, all four of the Tibes individuals exhibit an unexpectedly high reliance on dietary protein derived from freshwater fauna (river fish, shrimp, and crab) and terrestrial fauna (hutia, land crab, iguana), with smaller, but still significant, contributions coming from pinnipeds (seals), marine fish, and birds. In all four of the Tibes individuals, protein from freshwater and terrestrial sources accounts for at least 45 percent (summed means) of each individual’s total dietary protein. Third, for all four individuals, multisource mixture modeling reveals an unexpectedly low reliance on mollusks (both bivalves and gastropods) as a source of dietary protein. In all four individuals, mollusks account for, at most, 11.2 percent of dietary protein (summed means). That the Tibes individuals exhibited a large degree of dietary breadth is perhaps the least surprising finding of this attempt at modeling. Given Tibes’s po-

1st percentile Mean (%) 99th percentile

1st percentile Mean (%) 99th percentile

1st percentile Mean (%) 99th percentile

1st percentile Mean (%) 99th percentile

E-48

E-13

ES-3

CE-5

Individual Value

0.0 17.1 45.0

0.0 24.6 70.0

0.0 15.3 65.0

0.0 14.1 50.0

10.0 35.4 50.0

0.0 22.3 40.0

0.0 27.1 40.0

20.0 35.0 50.0

Freshwater Terrestrial Fauna Fauna

0.0 16.8 45.0

0.0 14.6 50.0

0.0 12.1 55.0

0.0 11.8 30.0

Birds

0.0 5.7 20.0

0.0 4.6 25.0

0.0 4.4 15.0

0.0 6.8 20.0

Turtles

0.0 2.5 10.0

0.0 3.5 15.0

0.0 3.2 10.0

0.0 3.8 10.0

Marine Fish

0.0 8.9 25.0

0.0 8.5 35.0

0.0 14.4 40.0

0.0 7.9 25.0

Marine Crabs

0.0 4.6 15.0

0.0 3.1 10.0

0.0 4.1 15.0

0.0 5.6 15.0

Gastropods

0.0 4.6 15.0

0.0 5.8 25.0

0.0 5.3 10.0

0.0 5.6 15.0

Bivalves

Table 10.5. First percentile, mean, and 99th percentile values for multisource mixture reconstruction of Tibes’s paleodiet

0.0 4.3 10.0

5.0 13.1 20.0

5.0 14.1 20.0

5.0 9.4 15.0

Pinnipeds

226 / William J. Pestle

Figure 10.2. Mean relative contributions of protein sources to diets of Tibes individuals.

sition in an ecotone between the coastal plains and the uplifted central highlands, many ecological niches would have been available for exploitation within more-or-less easy reach. Moreover, the unexpectedly high reliance on freshwater fauna suggested by IsoSource does not seem wholly unreasonable given the proximity of the site of Tibes to the Portugués River, from which such foodstuffs were likely derived. The size and apparent prehistoric abundance of the island’s freshwater shrimp (of which there are three genera, the largest of which can weigh upwards of 0.5 kg), fish, and crabs make this possibility seem all the more likely (deFrance et al., this volume). On the other hand, the relative unimportance of mollusks is still surprising, given that mollusk shells form such an enormous portion of the remains typically recovered from Puerto Rican archaeological sites, Tibes included. Such apparent ubiquity aside, the results of the present study suggest that mollusks would seem to have contributed only a small fraction (averaging less than 10 percent, reconstructed IsoSource mean values) of the protein portion of the diet of individuals interred at Tibes. This result does, however, mesh with both archaeological and ethnographic observations of a significant disconnect between the apparent observed quantity and the actual dietary contribution of mollusks in sites dominated by shell middens (e.g., Osborn 1977).

Bone Chemistry and Paleodiet at the Ceremonial Center of Tibes / 227 In terms of dietary protein, then, what is presented herein represents a significant departure from traditional archaeological reconstructions of the diet of ancient Puerto Ricans (e.g., Rainey 1940). Such reconstructions depicted the Puerto Rican paleodiet as focused primarily on the exploitation of marine crabs and mollusks, neither of which factor large in our reconstruction of the protein food web of Tibes. In fact, our research suggests the inverse—that more proximate freshwater and terrestrial resources formed the major portion of the protein diet of the individuals buried at Tibes, with marine and other sources playing a more minor role. This subsistence strategy meshes well with models of locally focused resource exploitation such as that suggested in Wing (1989). Turning to the energy portion of the human diet, the results are somewhat more straightforward, if still not wholly expected. As stated above, all four of the Tibes individuals possessed notably enriched d 13Cap signatures. Despite their diversity, plants fall into three isotopically distinct groupings as a consequence of their distinct photosynthetic pathways. In Puerto Rico, this can even be simplified further into just two groupings, as Puerto Rican CAM plants tend to appear, isotopically, like C4 plants. Consequently, in Puerto Rico, there are two distinct groups, C3 plants with a far more negative d 13C signature (averaging −25.05‰) and C4/CAM plants with a more enriched signature (averaging −9.52‰). Based on the enriched d 13C values derived from the apatite of the Tibes human samples and utilizing the formula provided above for the determination of percentage C4/ CAM in a given diet, the important role that C4/CAM plants must have played in the energy portion of the diet of all four Tibes individuals is evident (Table 10.3). From these calculations, C4/CAM plants were found to have accounted for an average of 44–78 percent of the dietary energy consumed by the Tibes individuals under analysis. It must be made explicit that in suggesting that a major portion of dietary energy was C4/CAM-derived we are not therefore implying that the specific source of dietary energy was maize, the most ubiquitous C4 cultigen in the New World. Macrobotanical evidence for maize from Puerto Rico is nonexistent, with En Bas Saline in Haiti and the Tutu site on St. Thomas being the nearest sites at which such evidence has been found (Newsom 1999; Newsom and Wing 2004; Pearsall 2002), and there are numerous other tropically distributed and edible C4 plants (e.g., amaranths and chenopods), the consumption of which may help to account for the calculated strong reliance on C4/CAM foodstuffs (Sage et al. 1999). That being said, in light of the recent recovery of maize starch grains from archaeological sites in Puerto Rico (Pagán Jiménez et al. 2005), the possibility that maize consumption may have played some significant role in the production of the observed enriched d 13Cap values of the Tibes consumers cannot be rejected out of hand. Given the temporal (more than two centuries at the limits of the 2s calibration) and possible sociocultural variation among the four individuals under

228 / William J. Pestle analysis, it is somewhat surprising that they form as tight an isotopic cluster as they apparently do. This apparent similarity does belie certain potentially significant isotopic and dietary differences among the individuals under study. While I am loathe to assign too much weight to differences observed among and between individuals in such a small sample, two aspects of these differences will nonetheless be briefly discussed here. First, the females from the Tibes sample, individuals E-48 and CE-5, differ substantially from the site’s males, E-13 and ES-3, in terms of the character of the protein portion of their diet. Both females possess markedly lower d 15N values and slightly more depleted d 13C collagen values than their male counterparts. These results would appear to suggest that the site’s female inhabitants consumed less seal and marine fish than their male counterparts and instead relied to a greater degree on terrestrial fauna for their dietary protein needs. While it is too early to draw any strong conclusions from these differences, they would appear to offer some suggestion of sex-based variation in diet among the members of the Tibes population. Similarly intriguing, and equally speculative, are the aberrant isotopic results for both the dietary protein and dietary energy of individual CE-5. This individual would seem to have been consuming the least high trophic level protein (as reflected in the lowest observed d 15N of the four individuals), the most terrestrial fauna and birds (as demonstrated by the most depleted d 13C collagen values), and the most C4/CAM energy (reflected in elevated d 13Cap values and the largest D13Cap−co spacing). While this apparent variation may be the result of the small sample under study or due to temporal differences (individual CE-5 is the earliest of the four individuals), it, as well as the differences in diet based on sex, may also be the result of exactly the sort of socially driven dietary inequality that the present study was conceived to examine. In fact, the exact type of dietary difference observed in individual CE-5, less protein and more C4/CAM energy, has been associated with low-social-status burials in other, admittedly geographically removed, complex societies (Ambrose et al. 2003). Taken as a whole, these preliminary results can perhaps also be used to provide some sort of insight into the broader nature of human activity at Tibes between a.d. 400 and a.d. 1200. Given that both the protein and energy portions of the diet, as we have reconstructed them here, deviate so notably from the faunal and botanical assemblages found at the site (as discussed by Newsom and deFrance et al., this volume), the possibility must be raised that the individuals buried at Tibes may not have been consuming all of their meals at the site. In fact, based on our results, it seems likely that only a small portion of the individuals’ meals (those focused on mollusks) were consumed and evidenced within Tibes and that a much larger portion (the freshwater fauna, terrestrial fauna, pinniped, and bird meals) were consumed elsewhere, at some removed domestic setting. This raises the pos-

Bone Chemistry and Paleodiet at the Ceremonial Center of Tibes / 229 sibility that the site was, in fact, a vacant ceremonial center and that the archaeological remains recovered there only reflect some small subset of the cultural practices of the people eventually buried therein.

Conclusions In sum, the present study both shows the promise of the application of stable isotope analysis to the resolution of paleodiet in Puerto Rico and provides some surprising insights into the diet of individuals interred at the Ceremonial Center of Tibes. The individuals buried at Tibes were consuming a substantial amount of high trophic level meat, as much as half of which was coming from traditionally underappreciated sources such as the nearby river and terrestrial ecosystems. In addition, they were supplementing their diets with marine elements (both fish and mammal) and birds, and they were only minimally reliant on more commonly suggested sources of food (crab and mollusks). The results of the study of the energy portion of their diets were similarly intriguing, as they indicated a large reliance on C4 or CAM plant materials, neither of which factor prominently in traditional reconstructions of the diet of the island’s prehistoric inhabitants. With these results in hand, I am extremely positive about the upcoming analysis of much larger samples from Tibes and three other precontact sites. Using stable isotope analysis of paleodiet, a high resolution and higher-fidelity technique than those traditionally employed in the reconstruction of paleodiet, I am extremely hopeful that greater insights into the nature of the site of Tibes and the development of in situ social complexity at that site can be better understood.

Acknowledgments This work was funded by a generous grant from the H. John Heinz III Foundation, part of the Heinz Family Philanthropies. In addition, this material is based upon work supported by the National Science Foundation under Grant No. BCS0612727. Finally, I wish to extend my gratitude to Antonio Curet, Edwin Crespo, Stanley Ambrose, the Instituto de Cultura Puertorriqueña, and the curators of the collections from Maruca, Tibes, Paso del Indio, and Punta Candelero.

Note 1. While the present chapter focuses on the largely encouraging results of the stable isotope analysis, it deserves to be noted that the other types of analysis employed in the pilot study, namely trace element and mtDNA, yielded what are, at best, mixed results. Specifically, trace element analysis revealed higher than acceptable

230 / William J. Pestle levels of diagenetic alteration of the mineral phase of the bone samples, although the tooth samples were found to be sufficiently unaltered by such processes. The mtDNA analytical methods employed were also found to be of limited efficacy, as they encountered numerous difficulties in the amplification of bona fide ancient DNA, with success in both amplification and sequencing achieved in only 1 of the 16 samples (Laffoon et al. 2006).

11 Tibes and the Social Landscape Integration, Interaction, and the Community Joshua M. Torres

Tibes represents an important place within the pre-colonial social landscape of the south-central coast of Puerto Rico. As a ceremonial center, the site was the focal point of social activity in which human agents produced and reproduced their society. Archaeological evidence from Tibes, based on stylistic characteristics of pottery (e.g., the presence of Hacienda Grande, Cuevas, and Elenan Ostionoid assemblages) in conjunction with the size and number of ceremonial features at the site, points to a long and complex history of use. However, it is important that our endeavor to understand Tibes not be limited to the boundaries of the site itself. To the contrary, Tibes was part of a broader network of people, places, and practices at a time when social, political, and ideological domains were undergoing dynamic transformations on the island of Puerto Rico and in the Caribbean in general. The south-central portion of the island, where Tibes is located, has a rich archaeological history. Over the past 70 years, field investigations in the region surrounding Tibes (e.g., Espenshade et al. 1987; Espenshade et al. 2007; Garrow et al. 1995; Rainey 1940; Robinson et al. 1985; Rouse 1952; Weaver et al. 1992; see also Rodríguez 1983, Appendixes I-A through I-D, for a listing of previous work) have supplied a corpus of data from which recent synthesis is providing a basis for understanding sociopolitical organization and development in the region (Curet et al. 2004; Torres 2001, 2005). In this chapter I continue to draw upon this body of archaeological data from the south-central coast to discuss the social landscape of which Tibes was a functioning part. To develop this discussion I present settlement and site-specific data to explore how the historical circumstances of settlement and geography shaped organization at local and regional scales around Tibes. This chapter posits that the devel-

232 / Joshua M. Torres opment of Tibes was in large part a product of the emergence of complex social networks and identity formation tied to the development of multivillage communities between a.d. 600 and a.d. 1200. Through this discussion, I also hope to provide further insights into the inception of post-Saladoid sociopolitical units in the region and contribute to recent issues regarding scale, social process, and archaeological interpretation in the Caribbean (Curet 2003; Curet et al. 2006; Curet et al. 2004; Keegan 2001, 2004).

Constituting the Community People do not exist in a vacuum—they relate to one another in a multiplicity of socially constructed ways and at a variety of interactive scales. To discuss social process at a meaningful scale above the individual site or village, it is critical to contextualize archaeological observations that address the ways in which individuals and groups interact and ultimately organize themselves within the physical and social spaces they inhabit. In recent research I have proposed that the community is a useful conceptual and methodological unit for approaching this issue (Torres 2005) because it provides a bridge between the local and the regional—a “focal point through which we can look at relationships on the local scale of analysis, as well as the interactions between communities, entering in and out of regional or global processes” (Carroll 1999:132). Central to this concept is the idea that the community and the village are not necessarily synonymous and the relationships that form the basis of social and political groups extend beyond the boundaries of a single spatial location or “site.” This supposition is founded on a consideration of the nonmutual exclusivity of various social relationships (e.g., kinship, corporate labor groups, gender) and the practices that facilitate social reproduction, specifically day-to-day interaction and ritual activity. These social relationships and the domains of social practice within which individuals and social collectives conduct their lives are the driving forces that structure and transform society (Bourdieu 1977; Giddens 1984). Importantly, these social relationships and domains of social practice have spatial and material correlates that can be tracked through time and provide us with clues to the organization of past social groups, as will be shown later in this chapter. Traditional perspectives regarding the community within archaeological and anthropological research contexts characterize it as a self-identifying residentially based territorial social group that interacts face to face on a consistent basis and is bound together in the shared utilization of local social and natural resources (Adler 2002; Kolb and Snead 1997; Murdock 1949; Varien 1999, 2002). As a function of these relationships, the spatial proximity of communities and the location of domestic occupation sites of their constituent members will influence the degree to which social groups interact and share forms of meaning and behavior

Tibes and the Social Landscape / 233 relative to their unique space/time contexts (Bourdieu 1977; Giddens 1984; Varien 1999; Yaeger 2000:125). This is not to say that communities are homogenous; to the contrary, the various social relationships that compose any social grouping allow for differences within and among its members. Therefore, a community represents a relational space that comes into being by virtue of the relationships between individuals that compose it as well as its relation to other communities within broader social networks (Cohen 1985). As spatial entities, communities are difficult to quantify despite the common archaeological practice to equate them synonymously with the “village” (Chang 1968:2–3; Hegmon 2002:265; Kolb and Snead 1997; Yaeger and Canuto 2000). There are many factors associated with the appropriateness of this terminological application; namely the scale and complexity of the site within the organizational contexts and the regional social system in which it exists (Hegmon 2002:265; Trigger 1978). However, since human relationships are a function of social, economic, political, religious, and ideological elements, individuals who form these networks will often crosscut single sites or villages. Further, the community will often spatially consist of multiple domestic groups that do not necessarily reside within the same location (Adler 2002:29). Important to the emergence and perpetuation of any community, and in particular political communities, is group-oriented ritual activity (Cohen 1985), and I would argue that all cohesive social groups that extend beyond consanguineal relations both rely on and are a product of practices that promote group solidarity and identity. The materiality of these practices is evinced in the built environment and in traditions of material culture. On the island of Puerto Rico archaeological evidence demonstrates that ritual activity was a critical component in the formation of social and political communities that led to the development of sociopolitical units, or cacicazgos, evident at the time of European contact (Curet 1996; Siegel 1991, 1996, 1999). From ethnohistoric documentation (e.g., Fernández de Oviedo 1975; Las Casas 1967), the material manifestation of group-oriented activity can be seen in the form of stone-enclosed ceremonial plazas and bateys that served as places for ritual dancing, feasting, and the playing of a ball game, with similar structural elements noted for Mesoamerican and South American societies. This concept is discussed in further detail in Chapter 12 of this volume. The material transformation between Saladoid and Ostionoid social space, from cleared areas, which were not physically defined, to spaces delimited by large stone rows or causeways, signifies a significant shift in the social structure on the island (Curet 1996; Siegel 1999). These public spaces present clear evidence of performative ritual practices that integrated individuals at multiple scales (Adler and Wilshusen 1990; Alegría 1983:5; Earle 1997:156; Hegmon 1989:5–14; Siegel 1999; Varien 1999:22). At these places, the collective social structure was reinforced through the practice of performative rituals in which community members en.

234 / Joshua M. Torres gaged in the intimate relation between ideology, power, identity, memory, and place (Inomata and Coben 2006). This material and social transformation is currently assumed to emerge on the south-central coast around the seventh century a.d. and is considered to be concomitant with the rise of self-aggrandizers who gained power through manipulation of ideology and ritual practices associated with ancestor veneration (Curet 1996; Siegel 1999). However, this perspective is inherently elite focused and negates other processes of social development; namely the ways in which communal identities are formed and how communities are both a medium and product of political consolidation (Pauketat 2007). Because of the time depth of material culture and materiality of performance evinced by the numerous plaza and batey features at Tibes, the site was clearly an important “place” within the social landscape. Based on the time depth and ideological importance of Tibes, its spatial, temporal, and material associations with other sites in the region implicate Tibes in a historical network of interrelationships that influenced processes of social organization and identity formation throughout the region. Supporting this supposition are observed shifting patterns in the materiality of the landscape on the south-central coast through time; namely changes in the morphology of settlement through time and space, similarities in the temporal rhythm in the construction of bateys/plazas at other proximally related sites on the south-central coast, and the transformation and distribution of pottery traditions.

Modeling the Social Landscape To examine the potential relationships between Tibes and other sites in the region it is necessary to create spatial and temporal frames of reference. In this chapter, I use topographic and archaeological site data within a geographical information system (GIS) to establish these contexts and explore the social landscape. Topographic data are in the form of a 1:20,000 digital elevation model (DEM) and were acquired from the U.S. Geological Survey (U.S. Geological Survey 2001). The DEM is a raster or grid-based elevation map in which each cell, or grid, in the map represents a 30-by-30-m area. Locational data for the majority of archaeological sites were acquired in 2001 in GIS format from the Officina Estatal de Conservacíon Historíca, and specific information regarding the sites (e.g., size, chronology, and potential function) was accumulated from a review of a variety of published and unpublished sources including cultural resource management contract reports, journal articles, books, and academic papers. Several sites were also documented as the result of a recent archaeological survey directed by the author in the region surrounding Tibes (Torres 2008). Information was entered into the GIS database

Tibes and the Social Landscape / 235 to manage and query the data (see Torres 2001, 2005 for more information regarding attribute structuring). Prior to discussing the data and how they are used in this study it is necessary to point out some of the problematic issues associated with the use of the data. The first issue is related to the nature of the sample of sites, which in many cases are not consistently identified as the result of systematic surveys. For instance, many sites have been located by accidental discovery (as in the case of Tibes) while others are a product of specific cultural resource management projects (e.g., Solís Magaña 1985; Thomas and Swanson 1986). Second, there are analytical problems stemming from variability in the registered information for sites as a product of site recordation strategies based on when and under what circumstances the work was conducted. Third, many sites have been destroyed by historical land practices and the expansion of urban development. Another important issue is time. The refinement of chronology in the area surrounding Tibes continues to be an important objective of archaeological research investigations in the region (e.g., Curet et al. 2006; Garrow et al. 1995; Krause 1989; Robinson et al. 1985; Rodríguez Ramos et al. 2010; Torres 2008; Weaver et al. 1992). Problematically, temporal assessment of a given site is often limited to the series or subseries level, which is based upon a now somewhat outdated sociotemporal framework for the island. Further, radiocarbon dating at two sigmas presents a range of variability that is in some cases up to 300 years (Keegan 2004; Rodríguez Ramos et al. 2010). In conjunction with the relatively short use life of domestic structures in the Caribbean of about 25 years (for further discussion see Bright 2003) and issues associated with their preservation, ascertaining contemporaneity between settlements is difficult at best. However, establishing general occupational sequences for sites through the use of artifacts and radiocarbon dates facilitates the examination of the social landscape with some degree of temporal resolution. As modifications to the traditional sociotemporal framework are currently in the process of development, I warily use the period classification devised by Irving Rouse (1992:107) with an emphasis on the time periods generally associated with the Saladoid (Period II, 400 b.c.– a.d. 600), the Elenan and Ostionan Ostionoid (Period III, a.d. 600–1200), and the Chican Ostionoid (Period IV, a.d. 1200–1500) pottery traditions. Since it has been demonstrated that these pottery traditions can deviate substantially from Rouse’s original temporal framework (cf. Rodríguez Ramos 2007; Rodríguez Ramos et al. 2010), I supplement the discussion with available radiocarbon dates registered from sites within the study area. Throughout this chapter, I will refer to mean dates calibrated to two sigmas. To represent the settlement pattern for the area around Tibes through time, sites have been divided into four types. However, rather than organizing or ranking

236 / Joshua M. Torres sites on an explicit system of presumed hierarchical relationships, I have categorized sites based on the potentiality for day-to-day and ritual activities. The categorizations are based on archaeological evidence of relatively long-term domestic occupation and the presence (or absence) of ceremonial features at any given site. Within this framework, sites were characterized as (1) sites with ceremonial features (plazas or bateys) and no domestic occupation, (2) villages with domestic occupation and ceremonial features, (3) villages with no ceremonial features, and (4) hamlet sites. Villages are typically defined in terms of evidence of relatively long-term occupation (based on ceramic assemblages and/or radiocarbon dates), household features, and/or substantial midden deposits. Differing from sites classified as villages, hamlets are in general smaller and, in many cases, represent a shorter time of domestic occupation (e.g., Espenshade 2000; Espenshade et al. 1987). Sites registered as camp sites and sites only possessing petroglyphs were not considered in this study, as emphasis is placed on sites possessing evidence of relatively long-term habitation. These sites are central to understanding the social landscape as the nexus of social practices and interactions. As a tool for examining the spatial relationships of sites surrounding Tibes through time, I generated a series of cost distance models around Tibes and documented settlements within the study area (cf. Torres 2005). Cost models measure surface distance based on impedance factors due to topography and have served in other archaeological research contexts as a proxy for traveling through the physical landscape and identifying areas of interaction between contemporaneous settlements (e.g., Torres 2005; Varien 1999). Differing from other distance-based models that rely on two-dimensional modeling techniques (e.g., drawing a 5-km circle around a site), cost models measure resistance units across a topographically nonuniform plane to calculate an accumulative distance surface from a given point location (see Wheatley and Gillings 2002:151–163 for detailed discussion). In this study, the U.S. Geological Survey DEM was used to generate slope that in turn was used to derive resistance values for travel. As in previous work (Torres 2005), the cost surface is generated using concepts developed by Cotterell and Kamminga (1990:195–196) and further quantified and employed by van Leusen (1999), which takes into account human travel costs (both up and downhill) for walking over the landscape. In this chapter I generated a cost model for 30 km as a visual tool for examining the distribution of sites surrounding Tibes. Further, I generated a 5-km cost model for settlements to show areas of potential interaction between settlements. The 30-km cost distance model, with Tibes as the center, defines the limit of investigation for the study (Figure 11.1) and establishes a boundary that approximates the maximal extent of a one-way day’s travel from the site. This distance is used for several reasons. First, this distance represents an inclusive range within which individuals at and around Tibes would have had the opportunity to interact with one

Figure 11.1. Map of Puerto Rico showing the project area.

238 / Joshua M. Torres another on a frequent basis. Second, since this distance represents a one-way day’s travel from Tibes, it delimits a physical constraint on land-based interaction and defines local spheres of interaction by those settlements in the immediate vicinity of Tibes and those located at greater distances (i.e., at the edges of the cost boundary farthest away from Tibes). This provides a local scale to examine site relationships in the immediate vicinity of Tibes, as well as a context for interactions at greater scales within the broader region. Finally, this area represents a conservative spatial realm within which residents of sites contemporaneous with Tibes would likely have had direct knowledge of Tibes and opportunities to visit the site and interact with members from surrounding settlements. I use a 5-km cost boundary for settlements to approximate areas most frequented by people as part of the practices associated with day-to-day activities within the landscape (Bourdieu 1977; Heidegger 1977; Ingold 1993). It is in the connections within these local spaces that people have the potential to intimately interact, form networks of social interaction, and develop shared forms of meaning and behavior necessary for the formation of communities that make up the basis of more complex sociopolitical units above individual settlements (Torres 2005). Within the study area a total of 107 archaeological sites have been previously documented; of these, 54 could be classified into one of the aforementioned site type categories and placed within a more specific temporal context based on relative (i.e., pottery) and/or absolute dating techniques (Table 11.1, Figure 11.2). Cost modeling was conducted for all sites designated as villages. This was done because, as mentioned previously, these sites form the nexus of local populations, the framework for community interrelationships, and the constituent component of sociopolitical groups. Based on the current information, 32 sites were classified as villages within the typological framework presented here.

Tibes and the Social Landscape Period II (ca. 400 b.c.–a.d. 600): Early Structural Developments In general, settlement locations during Period II (Figure 11.3) are situated along major river drainages on the coastal plains and in transitional areas adjacent to foothill locations (Curet 2005; Curet et al. 2004; Lundberg 1985; Torres 2001). This pattern supports previous research suggesting that early settlers were taking advantage of multiple ecological zones as part of opportunistic adaptive strategies related to local resources (Newsom and Wing 2004; Siegel 1991). In terms of colonization and the establishment of settlements, the logistical implications of settling in these areas would have not only maximized subsistence resource exploitation between inland and coastal settings but also facilitated travel along the coastal plain. As seen from the 30-km cost boundary, there is less cost in traveling

Sonadora

El Parking (PO-38) La Iglesia de Maraguez (PO-39) La Vaqueria La Mineral Los Gongolones Canas II Canas I Escuela Río Chiquito Reyes Ranchero Río Bayagan I Pico’s Ranchero Finca Feliciana Tecla

7

[8] [9]

Continued on the next page

10 11 12 13 14 15 16 17 18 19 [20]

Valle Verde Adjuntas 7 Vegas Arriba Zamas I Veguitas Zama I Mutaner

1 2 3 4 5 6

Map ID Site Name

0 0 0 0 0 0 0 0 0 0 1

0 0

0

0 0 0 0 0 0

HG

0 0 0 0 0 0 0 1 0 0 1

1 0

0

0 0 0 0 0 0

CU

Period II

1 1 1 1 1 1 1 1 1 1 0

1 1

0

0 0 0 0 0 0

0 1 1 0 1 0 1 1 1 1 1

1 1

1

0 0 0 0 0 0

OST

Period IIIa EL

Table 11.1. Site information within the project area

0 0 0 0 0 0 0 0 0 0 0

0 0

1

? ? 0 0 0 0

BOC

1 1 1 1 1 1 0 0 1 1 0

0 1

1

? ? 1 1 1 1

CAPA

Period IV

0 0 0 0 0 0 0 0 0 0 0

0 1

0

? ? 0 0 0 0

ESP

0 0 0 0 0 0 0 0 0 0 0

0 1

2

1 1 1 2 0 1

No. of Plazas

Torres 2008 Torres 2008 Torres 2008 Torres 2008 Torres 2008 Torres 2008 Torres 2008 Torres 2008 Torres 2008 Torres 2008 Siegel 1989

Site form Site form Site form; Alegría 1983 Site form; Alegría 1983 Site form Site form; Rodríguez Meléndez 2007 Site form; Rodríguez Meléndez 2007 Weaver et al. 1992 Garrow et al. 1995

Source

4 2 2 4 4 4 3 3 4 4 3

4 1

2

1 1 1 1 1 1

Site Types

La Vega Antes Cotui GU16 Santi Guayabal I Río Canas Venegas (Ollas Hondas) Collores

PE-1 La Jagua Tibes II Canas

PO-29 Duey/Diego Hernandez

La Fraternidad

Barinas II

[28]

29 30 31 32

[33] 34

35

36

21 22 23 24 25 26 27

Map ID Site Name

Table 11.1. Continued

0

0

0 1

0 0 0 1

1

0 0 0 0 0 0 0

HG

0

1

0 1

0 0 0 1

1

1 0 1 0 0 0 0

CU

Period II

0

0

1 0

0 0 0 0

1

1 0 0 0 0 0 0

EL

0

0

0 1

1 1 1 1

1

1 1 1 1 0 1 0

OST

Period IIIa

1

0

1 0

0 0 0 0

0

0 ? 0 0 ? 0 ?

BOC

0

0

1 1

0 0 0 0

0

0 ? 0 0 ? 0 ?

CAPA

Period IV

0

0

0 0

0 0 0 0

0

0 ? 0 0 ? 0 ?

ESP

0

0

2 1

1 0 0 0

0

0 0 0 0 0 1 2

No. of Plazas Site form; Siegel 1999 Site form Site form Site form Site form Site form Site form; Lundberg 1985:L16 Site form; Lundberg 1985; Pantel 1978; Rodríguez 1983; Rouse 1952 Site form Site form Pantel 1978 Site form; Chanlatte Baik 1976; Rainey 1940; Rouse 1952 Espenshade et al. 2007 Lundberg 1985; Maiz López and Questell Rodríguez 1990; Rouse 1952 Site form; Maiz López and Questell Rodríguez 1990 Site form; Maiz López and Questell Rodríguez 1990

Source

4

3

2 2

2 3 4 3

3

2 4 4 3 3 2 2

Site Types

Tibes

Hernandez Colon Lago Gely

Caracoles Las Ollas El Cayito

VL 4 PO-23 PO-27 Los Indios El Bronce PO-21

El Coco El Colmado Perez Las Flores

Cementerio De Guayanilla PR-10 midden

[38]

[39] 40

41 42 [43]

44 [45] [46] 47 [48] [49]

50 51 [52]

53

0

?

0 0 1

0 0 0 0 0 0

0 0 0

1 0

1

0

0

?

0 0 1

0 0 0 0 0 0

0 0 0

1 0

1

0

0

0

1 0 1

1 0 0 1 1 0

0 1 0

0 1

1

0

1

0

0 ? 1

1 1 0 0 1 1

1 0 0

1 0

1

0

0

0

0 ? 0

0 0 0 0 0 0

0 0 1

0 0

0

0

0

0

0 ? 0

0 0 1 0 1 0

0 0 0

0 0

0

1

0

0

0 ? 0

0 0 1 0 0 0

0 0 0

0 0

0

0

0

0

1 1 1

0 0 1 0 1 0

1 0 0

1 0

9

0

Torres 2008

Site form; Maiz López and Questell Rodríguez 1990 Site form; Curet et al. 2006; Siegel 1999 Maiz López 2002 Site form; Thomas and Swanson 1986 Site form; Rodríguez 1985 Site form; Rodríguez 1985 Site form; Garrow et al. 1995; Lundberg 1985 Site form Krause 1989 Krause 1989 Site form; Rodríguez 1985 Robinson et al. 1985 Espenshade et al. 1987; Espenshade et al. 2000 Site form Site form Site form; Aguilu cited in Wilson 1990; Siegel 1999 Site form 3

4

2 1 2

3 4 1 3 2 4

2 2 3

3 3

2

3

Note: Bracketed numbers represent sites for which there are registered radiocarbon dates. HG = Hacienda Grande; CU = Cuevas; EL = Elenan; OST = Ostiones; BOC = Boca Chica; CAPA = Capá; ESP = Esperanza. Site types: 1 = plaza/no domestic occupation; 2 = plaza/village; 3 = village/no plaza; 4 = hamlet. a Period III pottery designated to the level of subseries.

54

Y-9

37

Figure 11.2. Distribution of site types through time.

Figure 11.3. Settlements and cost boundaries, Period II. Bracketed numbers represent sites for which there are registered radiocarbon dates.

Tibes and the Social Landscape / 243 horizontally along the coast than vertically from the coast to the interior. This is a product of steeper terrain as one moves further inland vs. the relatively low relief of the coastal plain. Therefore, the ability to travel at less cost would have supported increased connectivity among social groups over greater distances along the coast. The ease of movement along the coastal plain, coupled with the potential for sea travel, may have influenced the orientation of regional social networks in a horizontal east–west fashion during this time. Site densities during this time are relatively low (n = 14) and there is no evidence supporting aggregation of settlements. At 5 km there is some clustering of settlements in the study area, with distinct groups in the west, east, and central portions of the study area (Curet et al. 2004; Torres 2005). Settlement and community organization during this time are thought to comprise large, relatively autonomous villages (Curet et al. 2004; Siegel 1999). Several sites in the region around Tibes, for instance Collores (Rodríguez 1983), Canas (Rainey 1940; Rouse 1952), and Tecla (Chanlatte Baik 1976), have yielded occupational evidence supporting this pattern. On the basis of data from these sites and others across the island from this time, there appears to be little variability in the spatial morphology (i.e., intrasite spatial organization) and general size of villages. The undifferentiated nature of villages, in terms of internal spatial organization, and the lack of obvious material correlates that indicate the formalization of status differentiation within or between sites suggest the development of sociopolitical units above the individual village level had not fully developed. Further supporting this are burials from this period at Tibes, and other sites on the island, in which preciosities and status items associated with formalized social inequality are absent (Curet and Oliver 1998; Siegel 1989). Within the study area immediately around Tibes, Saladoid pottery assemblages are characterized by the presence of Hacienda Grande (Period IIa, ca. 500 b.c.– a.d. 400) and Cuevas (Period IIb, ca. a.d. 400–600) traditions. Examination of the stylistic distributions of these traditions provides some clues regarding interaction in the area and offers a context for understanding the following period (Figure 11.4). Hacienda Grande pottery, located at several sites, and radiocarbon dates available for them indicate relatively early occupation of these villages (e.g., Tecla [Chanlatte Baik 1976]). Tibes also possesses Hacienda Grande pottery, although archaeological deposits yielding this material have not yet been dated (Curet et al. 2006). Observed patterning from sites possessing Hacienda Grande pottery shows their distribution throughout the study area and in relative proximity to the coastal plains, as mentioned previously. Toward the end of this period, sites with Cuevas and Hacienda Grande pottery extend further inland. This distribution supports the current idea of the expansion of settlements to interior areas toward the end of Period II or between about a.d. 400 and a.d. 700 in the south-central portion of the island (Curet 2005;

244 / Joshua M. Torres

Figure 11.4. Period II pottery stylistic distributions.

Curet et al. 2004; Lundberg 1985; Torres 2001, 2005). As sites during the later portion of Period II developed and expanded inland, networks of interaction would have consisted of not only horizontal connections along the coastal plains but also an increase in vertical connections from the coast to inland locations. Interestingly, Cuevas style pottery is present at many of the sites in the study area with Hacienda Grande style pottery. These sites are primarily clustered around Tibes but include Tecla, Canas, Collores, Diego Hernandez, Hernandez Colon, and Las Flores (Chanlatte Baik 1976; Curet et al. 2006; González Colón 1984; Maiz López 2002; Rodríguez 1983; Rouse 1952). The mutual presence of Cuevas and Hacienda Grande styles at these sites points to potential continuity in occupation sequences, transformations in material traditions, and historical interrelationships between these early sites on the south-central coast. Villages in the immediate vicinity of Tibes, particularly Hernandez Colon and Collores (Maiz López 2002; Rodríguez 1983), share similar stylistic sequences in pottery from Saladoid to Ostionoid, indicating a historical connection between the sites and, in a broader sense, the potentiality for a shared developmental trajectory above the scale of the individual village during this time.

Tibes and the Social Landscape / 245 Radiocarbon dates from sites within the study area that fall within Period II (400 b.c.–a.d. 600) provide a better context for establishing temporal relationships among sites documented with Saladoid components (Figure 11.5). Problematically, the sample of radiocarbon dates in the region during this time is limited to only three sites (Tecla, PO-23, and PO-38 [El Parking]) and the majority of the dates (n = 30) come from Hacienda Grande occupations at Tecla, a large village site located west of Tibes in Guyanilla (Chanlatte Baik 1976). Pottery in association with radiocarbon dates from the site show evidence of both Saladoid and Ostionoid components concurrently toward the end of Period II and prior to a.d. 600. Further, radiocarbon dates from the Cerrillos River valley, represented by sites PO-23 and PO-38 (Krause 1989; Weaver et al. 1992), also support the presence of early Ostiones pottery prior to a.d. 600. The relatively early emergence of Ostiones style pottery at Tecla, PO-23, and PO-38, in conjunction with the co-occurrence of Cuevas assemblages at these and other sites in the area, suggests a diversification in social practices and interactions above the level of the individual village beginning to develop in the region sometime around the fifth century a.d. Interestingly, most of the sites possessing both Hacienda Grande and Cuevas pottery occur in the cluster around Tibes and are visible within the central 5-km cost interval boundary immediately around the site. While very coarse, the distributional patterning points to areas of interaction or the potential for co-presence (Giddens 1984) of social groups, where multivillage community relationships were being developed through frequent interaction. Finally, and of particular note in relation to the radiocarbon dates, there is a potential temporal overlap for early Saladoid and Archaic populations in the region (Rodríguez Ramos 2001, 2005, 2007; Rodríguez Ramos et al. 2010). The presence of Archaic populations on the south-central coast has been documented at Cayo Cofresí and in Ponce at the site of Maruca (Rodríguez 1999; Veloz Maggiolo 1975). Radiocarbon dates from Cayo Cofresí (318 b.c. and a.d. 58 [Veloz Maggiolo 1975]) suggest potential temporal overlap between Archaic groups and Saladoid settlers on the south-central coast. This potential overlap has also been demonstrated in Loiza at the site of Maria La Cruz (Rouse and Alegría 1990). Undoubtedly, there was some interaction between these two groups that had some effect on settlement patterning and social development of later cultures in the region. However, further research is necessary to solidify the nature of these relationships and their subsequent impacts on community formation in the region. The general similarities in sites during this time in terms of settlement organization, pottery, and burial practices indicate some foundational structures that were well established prior to the sixth century a.d. These similarities in the material correlates between Saladoid social groups have been considered to be a veneer, a collection of superficial resemblances that served to unite small widely dispersed groups (Keegan 2000, 2004).

Figure 11.5. Calibrated radiocarbon dates (2-s ranges) for the south-central coast Period II.

Tibes and the Social Landscape / 247 I suggest that this Saladoid veneer was produced in the region around Tibes with a particular spatial emphasis on coastal settings due to accessibility of these places by land and water. This behavior would have mitigated risks involved in the colonization of the island and provided the presentation of a unified identity to social groups with divergent histories (e.g., Archaic peoples already living on the island and peoples of the La Hueca tradition). In contrast to the coastal plains, the foothill areas, where Tibes is located, make up a relatively secluded “back region” at the base of the foothills. This location represents a liminal space between geographical and perhaps social spaces where negotiations of power and identity were part of a recursive relationship between local and regional processes by the end of the period. I think that Tibes’s location in this area, at the edge of the foothills and coastal plains, contributed to its role as an important place in this regard. To conclude this portion of the discussion, settlement of the region by Saladoid pottery makers between ca. 400 b.c. and a.d. 600 represents a long and important period of establishing structural foundations and constraints for future development in the region. Tibes’s position in the landscape and the historical patterns of settlement that begin to emerge sometime between the fifth and seventh centuries a.d. set the stage in which intervillage communal ties at local levels became increasingly complicated and diverse. This is important because it is this historical context that leads to what archaeologists working in Puerto Rico consider the development of midrange societies and the formation of formal political institutions on the island (cf. Siegel 1999; Torres 2005).

Period III (ca. a.d. 600–1200): Diversification and Community Development Sometime after a.d. 600 a proliferation of new sites emerged throughout the south-central coast (Curet 2005; Curet et al. 2004; Lundberg 1985; Torres 2001, 2005) (Figure 11.6). Within the study area there is a dramatic increase in the number of settlements by approximately 160 percent (n = 37). Of the total number of villages from this time period, 12 sites (or approximately 35 percent) possess ceremonial architecture. Taking all documented sites in the region into consideration, there is a noticeable increase in the size and potential function of sites in general (Curet et al. 2004; Lundberg 1985; Rodríguez 1985; Torres 2001, 2005). In addition to changes in settlement patterning, several other changes show this period to be one of increasing social complexity and regional diversification. To begin with, this time is marked by the emergence of new pottery traditions that are more regionally variable than the previous Hacienda Grande or Cuevas traditions. Looking specifically at Tibes, the dynamic social changes that are occurring during this time are materially manifested in the development of formally delineated bateys and plazas, which appears to coincide with the peak in settle-

248 / Joshua M. Torres

Figure 11.6. Settlements and cost boundaries, Period III. Bracketed numbers represent sites for which there are registered radiocarbon dates.

ment of the region during Period IIIb, or sometime around the eleventh century a.d. (Curet et al. 2004; Lundberg 1985; Siegel 1999; Torres 2005). As sites increased in their frequency and distribution across the landscape, so must have interaction between denizens of the region (Curet 2005; Curet et al. 2004; Lundberg 1985; Torres 2005). This supposition is supported by observable decreases in distances between settlements. Increased interaction among contemporaneous villages, within the 5-km cost clusters, would have facilitated the extension of social and political networks outside of primary village contexts to incorporate members of multiple villages. This varies considerably from traditional conceptualizations of Saladoid sociopolitical organization in which settlements are considered to be relatively autonomous or “self-contained” political units (Siegel 1996), as previously mentioned. The recent identification of batey sites on the Portugués River, including PO-29, El Colmado Perez, and two new sites, La Mineral (PO-42) and Los Gongolones (PO-43), that were recently located and documented by the author (Torres 2008), shows that the immediate area surrounding Tibes was undergoing dynamic trans-

Tibes and the Social Landscape / 249 formations at this time. The presence of these ball courts, the variability in their size, and their relatively close proximity to one another are together reminiscent of observed community organization noted in Utuado for later periods (Oliver 1999). This is perhaps no coincidence, as the relatively open nature of the Portugués drainage extends to interior portions of the island and is less constricted than adjacent river valleys. This is important because this river drainage may have served as a highway to interior portions of the island, serving as a route for the movement of people and ideas between the coast and the interior. This idea is further supported by similarities in the structural configuration of Tibes and Caguana as well as by similar iconographic elements represented on petroglyphs at PO-29 and Caguana. This being the potential case, it is no surprise that Tibes emerged as an important center, as it occupies a space between the coast and the interior. Although the construction of the bateys/plazas at many sites adjacent to Tibes has not been formally dated, current archaeological data point to their development sometime between a.d. 600 and a.d. 1000 (Curet et al. 2006; González Colón 1984; Robinson et al. 1985; Rodríguez 1985; Torres 2008; Wilson 1991; also see Siegel 1999 for potential construction sequences). The ceremonial spaces constructed in the region are some of the earliest manifestations of these features on the island (e.g., Tibes and Las Flores). Interestingly, 50 percent (n = 7) of the sites possessing ceremonial architecture in the study area during this time are within the 5-km cost boundary from Tibes and five are in the Portugués River drainage. The intensification of the construction of ceremonial space throughout the island during this time, and in particular the south-central coast, suggests a widespread development and acceptance of social practices associated with ritual performance and the solidification of community groups. From a sociopolitical perspective, these sites have frequently been used to define centers of political power and as a primary indicator of hierarchical social relationships at a regional level (Oliver 1998; Siegel 1999; Torres 2005; Vescelius 1977). However, based on the variability in the size and space associated with ceremonial architecture (Siegel 1991, 2004), it would appear these features may have also served different purposes at different scales of social interaction (e.g., at the village or multivillage community level). For example, aside from Tibes, the four newly documented batey sites north of Tibes (PO-29, El Colmado Perez, PO-42, and PO-43) indicate local use of these spaces due to their relatively small size and inability to physically accommodate large numbers of people at any one time (especially in the case of El Colmado Perez, PO-42, and PO-43). Previous research, based on archaeological and ethnographic data, has shown a correlation between population and ceremonial spaces in which “[h]igh-level, ritually specialized facilities are typically utilized by large use groups, often including the members of several interacting, but separate communities” (Adler and

250 / Joshua M. Torres Wilshusen 1990:143). In the case of Tibes, the amount of space that developed between the eleventh and twelfth centuries, relative to other sites in the study area, points to the site’s ability to accommodate a relatively large number of people for ritual performance, supporting the notion that it was indeed a center at which social groups interacted at the supravillage level. Several villages registered with Saladoid and Ostionoid ceramic assemblages in the region surrounding Tibes, including Canas (Rainey 1940; Rouse 1952), Collores (Rodríguez 1983), and Tecla (Chanlatte Baik 1976), all possess evidence of long-term occupation but do not show evidence for the development of ceremonial architecture. Why is this? Did it never develop at these sites or have formation processes and time erased any evidence of their existence? Assuming that no ceremonial architecture was built at these sites, we must ask, “Why not?” Indeed if, as Siegel (1999) has suggested, ancestor veneration was an indicator of centralized power and the territorialization of social spaces, then why is this not evident at these sites, which possess both occupational depth and, by virtue of this, vested interest in their place within a changing social landscape? One would expect the sites in relatively immediate spatial and temporal proximity to Tibes to be subject to similar historical contingencies related to territorialization and community development. Since monumental features did not develop at these locations, then it seems logical to conclude that social processes were contextually contingent on local developments within a broader regional system, as previously suggested by Curet (2003, 2005; Curet et al. 2006; Curet et al. 2004). Assuming that pottery traditions, at some level, represent material manifestations of social identities coincident with historical networks of interaction, examination of pottery distributions during this period offers further insight into the potential relationships between people and place in the area around Tibes and the region in general (Figure 11.7). Typically, pottery styles of the Ostionoid series are considered to be more or less affiliated with social groups from the Mona Passage (Ostionan) and Vieques Sound (Elenan) areas, whereby the spatial distributions of pottery styles are found in varying ratios between the east and west portions of the island and appear to be related to the distance from either area (Goodwin and Walker 1975; Robinson et al. 1985; Rouse 1952). In the area around Tibes, it has been noted that many sites show evidence of Elenan and Ostionan assemblages from associated stratigraphic contexts mixed in similar proportions (Robinson et al. 1985; Rodríguez 1985; Thomas and Swanson 1986; Weaver et al. 1992). And while this mixing of assemblages is not uncommon in different parts of the island (e.g., Goodwin and Walker 1975), the frequency of occurrence and similarity in ratios at some sites around Tibes suggest that the region was a point of articulation in which social groups historically associated with the Mona Passage and the Vieques Sound areas overlapped. The differentiation of east and west horizontal interaction spheres manifested

Tibes and the Social Landscape / 251

Figure 11.7. Period III pottery stylistic distributions.

in Ostionoid styles begins to emerge in this area around the fifth century a.d. with Ostiones style registered just east of Tibes in the Cerrillos River valley at sites PO-23 (14C 2s mean a.d. 427) and PO-38 (14C 2s mean a.d. 598). Shortly after the seventh century a.d., several sites around Tibes begin to show mixing of pottery styles at sites like El Bronce, Lago Gely, Collores, and El Parking (Robinson et al. 1985; Rodríguez 1983; Thomas and Swanson 1986; Weaver et al. 1992). Currently, the earliest solid evidence of mixed Ostionoid style pottery comes from El Bronce (with Ostiones and Santa Elena pottery), to the southeast of Tibes (14C 2s mean a.d. 753 [Robinson et al. 1985]). In addition to the mixed pottery assemblages at sites around Tibes, some sites in the area do actually display more affiliation with Ostiones or Elenan styles. For instance, Hernandez Colon, a site within 5 km of Tibes and with evidence of earlier Saladoid occupation, is primarily characterized by Ostiones ceramics (Maiz López 2002), as is PO-21 in the Cerrillos Valley east of Tibes (Espenshade et al. 1987), suggesting stronger ties to western Puerto Rico. Conversely, Tibes, PO-42, PO-43, and PO-29 all display more affiliation with Elenan styles, which would indicate stronger ties to areas in eastern Puerto Rico. Within the study area during

252 / Joshua M. Torres this time, sites possessing similar pottery assemblages are often immediately adjacent to one another. Examination of the distribution of pottery styles by site clusters throughout the study area generally shows conformance to an expected pattern whereby sites in the west are more homogeneous—represented primarily by pottery of the Ostiones subseries—and sites in the eastern portion of the study possess more Elenan subseries. However, in the center of the study region immediately surrounding Tibes, there is considerably more diversity in the representation of pottery styles, suggesting that the social groups settled around Tibes, while developing locally, had historical connections to groups in eastern and western Puerto Rico. Further, it points to the area around Tibes as being part of a larger system of interaction between social networks from these two areas and as a point of articulation of people and ideas from different spheres of historical and social influence. Radiocarbon dates registered from several sites during this period provide a context to further consider settlement history and interaction. The number of sites with radiocarbon dates is substantially higher than that for Period II, with 42 dates registered from 12 sites (Figure 11.8). Overall, five radiocarbon dates from this period are currently registered for Tibes (2s mean a.d. 822, a.d. 947, a.d. 1020, a.d. 1097, and a.d. 1114 [Curet et al. 2006]). These dates, in comparison with those from other dated sites in the region, suggest it is highly probable that many of these sites were contemporaneous. Just by looking at sites within the 5-km cost boundary around Tibes, several sites immediately stand out in this regard. In relation to the early radiocarbon date at Tibes (2s mean a.d. 822), El Bronce (2s mean a.d. 830) and Collores (2s mean a.d. 825) show the closest affiliations. This is not surprising, as these sites also possess significant amounts of Elenan pottery. The radiocarbon dates also point to several sites outside the immediate vicinity of Tibes that may be contemporaneous. Looking at the early date for Tibes (a.d. 822), the sites Tecla (2s mean a.d. 817) and Las Flores (2s mean a.d. 842) are extremely close temporally. Notably, Las Flores displays evidence for Elenan assemblages, whereas Tecla is primarily characterized by Ostionan assemblages. And while all of these sites possess evidence of being occupied during Period II, the potential connections between Tibes and Las Flores seem to be stronger based on the development of ceremonial architecture and similar pottery traditions at these two sites. Later dates for Tibes support some of these relationships and further support potential contemporaneity with several other sites in the region. For instance, one date from Tibes from the beginning of the eleventh century a.d. (2s mean a.d. 1020) is very close to three other dates from Las Flores (2s mean a.d. 1000, a.d. 1030, and a.d. 1035). Further, two sites from the Cerrillos River valley, PO-39 (2s mean a.d. 998) and PO-38 (2s mean a.d. 1049), also display close temporal

Figure 11.8. Calibrated radiocarbon dates (2-s ranges) for the south-central coast Period III.

254 / Joshua M. Torres proximity to this date. The connections made through the associations of radiocarbon dates point to several sites in the region around Tibes that were potentially interacting and subject to similar social processes of change and development at the regional level. As networks along the coast continued through time, east–west differences were also likely influenced by interior-to-coast relationships that were concomitantly developing with patterns of population growth and dispersal during this time. The connections between coastal and inland settlements are evident at Tibes and other sites immediately north of it where substantial quantities of marine fauna are located in midden deposits (Curet et al. 2006; Torres 2008; see also deFrance et al., this volume). Based on the available data, it appears that social networks during this time were evolving in a context of differing social realms of interest (Garrow et al. 1995:233) or regional differentiation characterized by east–west and interior-to-coast interactions. These social networks, built on daily social practices and ritual activity, set the rhythm and tempo of the formation of new and increasingly complex frames of social reference during this time. I would agree with Siegel that as community networks became increasingly complex, sociopolitical competition increased (Siegel 1999). However, based on the variations in size and number of ceremonial features it also seems that ceremonial architecture served varying functions at different scales of social organization rather than strictly demonstrating hierarchical organizational relationships across the landscape. From this perspective Tibes would have served functions at multiple scales for both the local community of which it was a part and the multiple communities that surrounded it. This time represents a period of obvious diversification in material tradition and the potential for the emergence of new social and political identities developed through historically contingent interaction spheres in different parts of the island. This is quite different from the Saladoid “veneer” and similarities in material culture noted in the region from the previous period. Importantly, this was a period of increasing sociopolitical complexity characterized by the development of multivillage community networks (Torres 2005) focused on the integration of people. These communities were tied together by ritual behaviors in which differences between groups were abolished through performance-oriented ritual activity that facilitated social solidarity in the context of increasing regional diversity.

Period IV (ca. a.d. 1200–1500): Abandonment and Demographic Change Sometime after the thirteenth century a.d., there is an apparent abandonment of many settlements (including Tibes) in the region (Figure 11.9). During this time site frequencies decrease by approximately 33 percent from 37 to 25 settlements

Tibes and the Social Landscape / 255

Figure 11.9. Settlements and cost boundaries for Period IV. Bracketed numbers represent sites for which there are registered radiocarbon dates.

within the study area. This reduction of sites has been suggested to be a result of a demographic shift in which populations may have aggregated to settlements back down on the coast (Walker, personal communication 2005). However, this population shift may also have been the result of other factors such as hurricane activity (Rodríguez 1985) or the cyclical nature of incipient political formations in which social groups fission and fusion in different stages of the political cycle (Anderson 1996a, 1996b). Regardless, the observed shift represents a significant change in the regional organization of populations (Curet 2005; Curet et al. 2004). Interestingly, several settlements immediately north of Tibes in the Portugués River (PO-29, PO-42, PO-43, and El Colmado Perez) all point to primary occupations from after a.d. 600 and into the Chican Ostionoid period (Espenshade, personal communication 2008; Torres 2008). Several of these settlements are poorly understood and the disuse of Tibes and other major village sites in the region represents some of the most interesting and perhaps important avenues of archaeological inquiry yet to be addressed for the south-central coast. However, the presence of these sites with ceremonial features and their continued use past ap-

256 / Joshua M. Torres proximately a.d. 1200 suggest continued connections between the coast and interior portions of the island in which the Portugués River may have served as a highway from the coast to the interior. The decrease in sites along the southern coast is accompanied by a proliferation of sites in the interior portion of the island. As displayed in Figure 11.9, seven sites are located in these upland areas and are probably more related to upland settlements than southern settlements based on distance and ability to travel through these areas. The gap between these sites and those to the south may be due to a lack of sites in the intervening areas but also may be a product of the lack of archaeological survey work conducted in these areas. The relationships between the upland mountains and coastal settings have not been well studied in this region in general. Based on available information regarding these sites, several appear to be ceremonial with no or limited evidence of domestic occupations (Alegría 1983). This would coincide with the pattern observed for sites in Utuado during this time in which ceremonial sites lacking long-term domestic occupation are more common than in the preceding period. However, it has been noted that there are many village sites in interior settings that also possess ceremonial architecture (Oliver 1998). The presence of ceremonial centers with no formal domestic occupation is an interesting development and is seen on the south-central coast, during Period III, at PO-39 (Garrow et al. 1995). This development and the function of these sites within the settlement system are also not clearly understood. Vescelius (1977) has noted that ceremonial centers may lie on the border of political units and serve to integrate dispersed social groups. From this perspective, unoccupied ceremonial sites could be gathering places that served multiple communities from adjacent areas. Data related to the distribution of pottery styles during this time are somewhat limited: one-third of the sites in the study area (n = 11) have no information at the level of style. The stylistic information that does exist is interesting, and in general some patterning can be discerned. Capá style pottery, typically associated with western Puerto Rico, is found in low frequency throughout the study area, with a small cluster of sites with Capá in the northern portion of the project area. Esperanza pottery, typically associated with eastern Puerto Rico, appears in only one registered site in the far eastern portion of the study area (Figure 11.10). Two sites in the Cerrillos River valley, PO-39 and PO-27, both contain mixed Capá and Esperanza assemblages, which suggests some connectivity in east–west relationships in the area. One other interesting observation is the presence of Boca Chica style ceramics in the area at El Cayito, Sonadora, Barinas II, and potentially PO-29 (Espenshade et al. 2007; Lundberg 1985). These ceramics, typically associated with the eastern Dominican Republic, appear to be both in the upland areas represented in this study and at sites along the coast. Sites possessing this pottery

Tibes and the Social Landscape / 257

Figure 11.10. Period IV pottery stylistic distributions.

have been explained as either trade related or possibly outpost settlements from the Dominican Republic (Lundberg 1985). Radiocarbon dates from this time are very limited, with five dates registered from the following sites: El Bronce (2s mean a.d. 1268), El Cayito (2s mean a.d. 1295), Tibes (2s mean a.d. 1290), Las Flores (2s mean a.d. 1352), and PO-27 (2s mean a.d. 1369). The date from Tibes is the latest currently registered for this site and was found in association with Santa Elena pottery. The late dates for El Bronce and Las Flores show the mixed Ostionan and Elenan assemblages (Robinson et al. 1985; Wilson 1991). Finally, the dates associated with El Cayito and PO-27 are related to Boca Chica and mixed Capá and Esperanza pottery, respectively. Based on these data, and assuming the dates are correct, it would seem that Tibes and Las Flores, ceremonial centers from the previous period, retained the use of older styles. Further, the mixed assemblages at PO-27 suggest that social networks, although regionalized, were connected even with relatively small sites like those of the Cerrillos River valley. Finally, the dates indicate that many of these sites were occupied well after the thirteenth century. To conclude this portion of the chapter, Period IV represents a time of major

258 / Joshua M. Torres structural transformations that appear completely different from and perhaps counter to the presupposed evolutionary trajectory proposed for the previous period. The dates from Tibes and some other sites in the area (e.g., El Bronce) and the Chican assemblages associated with batey sites along the Portugués River suggest that the “peaceful collapse” of Tibes appears to occur sometime after the fourteenth century. Regardless, the shift in settlement throughout the region between the thirteenth and fifteenth centuries certainly resulted in a dramatic transformation in the organization of local and regional social networks on the south coast. Within these changing networks it is quite possible that Tibes lost its importance due to an inability to adapt to these newly developed social networks (Curet 2005; Curet et al. 2004; Torres 2001, 2005). However, the exact reasons that Tibes appears to have fallen into relative disuse are still, and perhaps always will be, a mystery.

Concluding Remarks The development of Tibes as an important ceremonial center on the south-central coast is a product of several complex processes. One of the main processes associated with the development of Tibes was the relationships developed between people throughout the landscape as part of day-to-day and ritual social practices, which influenced the degree to which groups shared forms of meaning and behavior as individuals and as members of a community (Bourdieu 1977; Varien 1999; Yaeger and Canuto 2000). These networks of interaction surrounding Tibes manifested themselves during the Saladoid along the coastal plains where people, material, and information could be transported relatively easily over land or by water. By the end of this period, transitions in stylistic characteristics in pottery denote an increase in the complexity of social networks in the region and diversification of social identities. Based on mixing of pottery styles at many sites in the region between the seventh and thirteenth centuries a.d., the area surrounding Tibes was a place of articulation between groups rather than an area of separation. Importantly, by the following period the Portugués River appears to have become an important route from the coast to the interior, exemplifying the development of interior-to-coast networks during this time. Examination of settlement patterns and characteristics of pottery styles for sites surrounding Tibes through time suggests the emergence of complex multivillage social networks sometime around the seventh century a.d. The development of these networks can be seen as a product of historically contingent processes related to settlement location and tied to the veneer of Saladoid tradition that was later transformed through contestations of place and power, as noted by Siegel (1996, 1999). Further, I also suggest that these transformations in the negotiation of social identities were tied to the historical contingencies of place in a dynamic landscape.

Tibes and the Social Landscape / 259 In consideration of Tibes, day-to-day and ritual practices conducted within and between villages surrounding the site formed part of a recursive relationship between the quotidian (foreground) and idealized (background) potentialities related to the process of place and social structuration (Giddens 1984; Hirsch 1995). As ritual performance forms the foundation of communal relationships, the materialization of ceremonial features evident at Tibes and other sites in its immediate proximity, particularly in the Portugués River drainage, integrated groups of people at varying scales. It is no surprise that the construction and use of formalized ritual spaces became accepted practice in the region surrounding Tibes. The confluence of various social groups in the region surrounding Tibes offers another dimension for examining the emergence of early political units in which social complexity emerges from what has been termed “complex socialities” (Sassaman and Randall 2007). Central to this concept is the emergence of forms of alterity and mimesis in the negotiation of interaction between people with different identities. From this perspective we can begin to focus less on the strategies used by self-aggrandizing elites to subjugate the masses and more on how people structured their social realities at local and regional scales. It is in these contexts that communities develop, they become politicized, leaders emerge, and incipient polities develop (cf. Pauketat 2007). I realize that the interpretations presented here are tentative; however, I feel that understanding Tibes and the people who created the social landscape during its heyday requires making connections between people, places, and time. The complex nature of Tibes and its importance on the south-central coast stem from its implied role as a place of interaction between communities. Future research regarding Tibes and its role in the development and organization of the social landscape of the south-central coast should be tempered by archaeological inquiry related to refining regional chronology and understanding how social groups were organized and interrelated at varying scales. This can be accomplished by focusing on the comparative intersite artifact analyses between proximally related, contemporaneous sites to establish variability in particular artifact types that would give clues to interrelationships between sites and indicators of power and identity through material culture. Additional research at Tibes and on the south-central coast will, it is hoped, elucidate some of these complex processes in the future and provide a firmer basis for the interpretation of the ideas presented in this chapter.

Acknowledgments I would like to thank Antonio Curet for offering me the opportunity to participate in this exciting project and for his continued help and support of my research. Special thanks to Lisa Stringer for assisting on the many edits of this contribution.

260 / Joshua M. Torres Thanks to my colleague and friend Reniel Rodríguez Ramos for his advice and input on several aspects of this chapter. He provided me with many of the dates for the sites mentioned, which he has painstakingly compiled over the past few years from various sources. Finally, I would like to thank two anonymous reviewers for commenting on an earlier draft of this chapter. Their insightful comments greatly added to my thinking and to reorganization of the chapter. I take sole responsibility for any shortcomings or problems associated with this chapter.

12 Plazas, Bateys, and Ceremonial Centers The Social and Cultural Context of Tibes in the Ancient History of Puerto Rico L. Antonio Curet and Joshua M. Torres

The presence of ceremonial centers in the Greater Antilles has always been the source of much debate, speculation, and imagination in Caribbean archaeology. Throughout history they have been indiscriminately used as evidence for cultural affiliation (Rouse 1948), sociopolitical development (Curet 1992a; Oliver 1998, 2009; Siegel 1999), ideological structures (Oliver 1998, 2009), social interaction (García Arevalo 1990; Oliver 1998; Siegel 1996, 1999), and diffusion from the so-called greater civilizations of Mesoamerica (Alegría 1983; Fernández Méndez 1972; García Goyco 1983). However, with the exception of Oliver’s and Rivera Fontán’s work in Caguana and its surrounding region (Oliver 1998, 2009; Oliver et al. 1999; Rivera Fontán 1992, 1999; Rivera Fontán and Oliver 2005), that of Rivera Fontán and Silva (1997, 2002) in Mayaguez, and that of Rodríguez Meléndez (2007) in Jayuya, most of the interpretations of the function of these structures and their development rely almost entirely on historical descriptions or superfluous (or superficial) comparisons with other regions. Furthermore, these structures and types of sites are usually considered “autonomous” phenomena that operated in a social and cultural vacuum, without considering the context (i.e., regional or extraregional) of which they were an essential part. It is our contention that in order to understand the role and function of ceremonial structures and, more important, ceremonial centers, we need to gather and consider more detailed archaeological data. This approach will allow us to gain a better understanding of the social, economic, political, and historical context in which these sites operated. The purpose of this chapter is to discuss some of these issues based on available evidence for the civic-ceremonial center of Tibes. Tibes is ideal for the study of the

262 / L. Antonio Curet and Joshua M. Torres development and function of ceremonial centers in the ancient societies of Puerto Rico, first, because it is one of the earliest examples of such sites in the region and, second, because of its long history (see discussion below). We have evidence of the early humble beginnings of the site as the habitation location of a nonstratified group to the major development of the ceremonial center that most probably involved extensive and deep changes in social and political organization. In the sections that follow, we first discuss how ball courts, plazas, and ceremonial centers have been described and interpreted and the primary issues surrounding these features that have been debated by Caribbean archaeologists. This discussion is followed by a synthesis of the information gathered for Tibes and presented in this volume by the various authors of the previous chapters. We end with a discussion of how this evidence contributes to our understanding of the function and role of the structures and the ceremonial center in the social and cultural histories of ancient Puerto Rico. Although we are not ready to answer many of the questions related to the development of these sites and the implied social changes associated with their construction, the results from this study facilitate our understanding and help contextualize the role of these sites in the political landscapes before European contact.

Ball Courts, Plazas, and Ceremonial Centers Terms and Definitions Prior to discussing ceremonial structures and centers it is important to present some definitions that will lay the conceptual foundation for this discussion. First, we follow Oliver’s (1998) terminology for naming the different types of structures. Based on their shapes, three types of ceremonial structures are recognized in the Greater Antilles: rectangular, quasi-quadrangular, and circular. Because of the nature of the ball game and the description in the early European documents (see below), it is widely accepted that rectangular structures were likely used for the ball game. The native people of the Greater Antilles used the Taíno word batey to refer to the ball game, the ball court, and the ball itself (Las Casas 1967:II:350). In this work, we use the terms ball court and batey interchangeably when referring to these rectangular structures. Because of their square shape, quasi-quadrangular structures do not seem appropriate for playing a ball game but ideal for communal ceremonies, dances, and other ritual activities. For this reason, we use the term plazas. Similarly, round structures do not seem functional as ball courts and they, too, are referred to here as plazas, while keeping in mind that differences in shapes, compared to the quasi-quadrangular structures, may also reflect differences in function, use, and meaning. Beside these three shapes, other sites include smaller and, in most cases, marginal structures, for example, cobblestone causeways or stone alignments that look like possible retention walls.

Plazas, Bateys, and Ceremonial Centers / 263 A cautionary word on this terminology is necessary here. This classificatory system and its terminology suffer from the perennial problem archaeologists face when relating form and function in developing typologies of artifacts, features, or sites. Although in many cases the relationship between form and function is obvious, it is not necessarily true in all situations since one form can be related to more than one function or some forms may have been “misused” with an unrelated function (e.g., a screwdriver used as a hammer or a chisel). This is true in the case of the ceremonial structures in Puerto Rico. Some “ball courts” may have been used for different or multiple functions (e.g., dances or other ritual ceremonies) and ball games may have been played in quasi-quadrangular plazas or even outside the structures. In fact, some of the structures named ball courts in this volume may have been too small for playing the ball game and they may have been more symbolic than functional. Therefore, the terms mentioned above are used in this chapter as terms to discuss the types of structures described, keeping in mind that they may have had different or multiple uses and meanings. It is also important to mention that throughout the Greater Antilles, the structures vary not only in shape but also in size and construction materials. Sizes vary from 1,761 m2 in the case of some of the largest ball courts in Puerto Rico to 125,016 m2 for the circular plazas from the Dominican Republic. There are two main types of construction material for the structures: stone and soil. As discussed below, the chronicles mention that ball courts in Hispaniola were defined by earthen berms. Examples of this kind of construction are evinced by several structures in Cuba (Alegría 1983; Tabío and Rey 1966), although they have also been reported in Puerto Rico (González Colón 1984; Rouse 1952:484–488). In Puerto Rico, the material used par excellence in the construction of plazas and ball courts is stone. Structures with stone rows similar to those found in Puerto Rico can be found in eastern Hispaniola, but large circular plazas bounded by stones and ball courts defined by earthen embankments are primarily found in other parts of the latter island (Alegría 1983). The geographic distribution of these structures also varies considerably throughout the Greater Antilles and the Caribbean (see Alegría 1983; Wilson 1990). Puerto Rico has the highest documented frequency by far, and possibly the highest density, of structures in the Caribbean archipelago. Densities in Hispaniola are relatively low and tend to be concentrated on the eastern side of the island. However, round structures, larger than any Puerto Rican plaza or ball court, are found in the central part of Hispaniola. Similarly, Cuba has very few structures, mostly bounded by earthen embankments, and they are concentrated on the eastern part of the island. Some of them are considerably larger than the Puerto Rican examples (Guarch Delmonte 1972; Tabío and Rey 1966). Other terms that need some discussion in this section are those used for the concepts of “ceremonial center” and “civic-ceremonial center.” In Puerto Rico,

264 / L. Antonio Curet and Joshua M. Torres these terms usually refer to a site with multiple ball court/plaza structures. From this perspective, sites with one or two structures are not typically referred to as “centers.” However, in other islands such as Hispaniola and Cuba the definition associated with these terms may have different connotations. In both of these cases, because of the low density of ceremonial structures, a few sites with only one large structure (in some cases considerably larger than the Puerto Rican structures) can be considered ceremonial centers, especially when the size of some of the circular plazas is considered. Thus, the definition of what a ceremonial center is can vary from island to island. Furthermore, some of us have been using the term civic-ceremonial center to imply that the site may have been not only a religious center but also a center of social and political power. However, the application of this term is based on our assumptions of site use, which relate back to the aforementioned concepts of form and function, which do not always coincide and are aspects that need to be further substantiated with empirical data. In this chapter, we exclusively use the term ceremonial center to refer to the Puerto Rican multistructure sites in order to avoid problems associated with the concept of civicceremonial centers. These concepts and our use of them are not static and we anticipate that future archaeological work will produce information that may change our definitions of both terms. For example, the possibility exists for the presence of ceremonial centers without ball courts and plazas but with large buildings used in communal rites. Ceremonial centers of this type have been reported ethnographically for several regions of South America, where a communal, ceremonial building is present in a centralized village and serves settlements in the surrounding region (e.g., Thomas 1982).

Ball Courts and Plazas The first reference to ceremonial structures in the Greater Antilles was presented by early European writers when they mentioned the ball game played on Hispaniola (e.g., Fernández de Oviedo 1959b; Las Casas 1967:II:350). While most of these writers mention the presence of specialized “plazas” used for the ball games, Las Casas (1967:II:350) is the only chronicler who actually describes them. His description indicates that ball courts were characterized by their length being three times greater than their width (i.e., rectangular) and by their being bounded by earthen berms 20–40 cm high (“un palmo o dos de alto”). These dimensions are often applied as standard measurements for the identification of ball courts in archaeological sites. In the history of Puerto Rican archaeology, Jesse Fewkes (1970) was one of the first scholars to report (in 1907) the presence of plazas or ball courts. However, it was not until the early field investigations of J. Alden Mason in 1915 at the site of Caguana that these structures were given serious consideration (Mason 1941). Caguana was the first ceremonial center identified and studied systemati-

Plazas, Bateys, and Ceremonial Centers / 265 cally in Puerto Rico. Although earlier pottery is found at the location, most of the site appears to belong to the late pre-Hispanic period. Mason’s meticulous work and reporting of the site in which he described the structures, petroglyphs, and buried features is one of the jewels of Puerto Rican archaeology. Mason excavated and mapped the structures and unearthed evidence suggesting that the site had a long history of construction and reconstruction of buildings and ceremonial structures. More recent work at this site (Rivera Fontán 1992, 1999) and in the region (Oliver 1998; Oliver et al. 1999; Rivera Fontán and Oliver 2005) demonstrates that this ceremonial center is only one part of an intricate regional system, wherein physiography, social organization, and economics intertwined to create a dispersed settlement pattern of small hamlets, some of which include their own ball courts (Oliver et al. 1999). Later researchers such as Rainey (1940:98–102) and Rouse (1952) reported structures in a number of sites, but little attempt was made to study them in the larger regional context in which they existed. In 1983, Alegría published his nowclassic book on ball courts and plazas in Puerto Rico and the Caribbean, in which he surveyed the information available on sites with ceremonial structures. Alegría points out differences and similarities between structures known at that time from the different Greater Antilles. He also presents a short report of his excavations and the restoration of Caguana in the late 1940s, suggesting that this site may have actually been a vacant center (Alegría 1983:87). Another conclusion reached by Alegría is that the ball game and the ball courts were probably the result of Mesoamerican influence that diffused through lower Central America, then northeastern South America, eventually reaching the Greater Antilles. Other researchers such as Fernández Méndez (1972) and García Goyco (1983) also reached a similar conclusion, but in these cases they claim the diffusion flowed directly from Mesoamerica to the Greater Antilles. However, Walker’s (1993) study on the iconography of ceremonial objects related to ceremonial structures has questioned these conclusions and, instead, argues that they seem to have a South American origin. More recently, Rodríguez Ramos and Pagán Jiménez (2006, 2007) and Wilson (2007) have pointed out similarities in the use of causeways in some Puerto Rican structures and the ones found in Costa Rica. They both have suggested a possible connection between the two regions that included the transmission of ideas, symbols, beliefs, and architectural meaning. The initial studies at the Ceremonial Center of Tibes (as discussed in Alvarado Zayas and Curet, this volume) by the Sociedad Guaynía uncovered two primary occupational sequences. The first one, containing a Saladoid component, seems to have been a village, and the second one, possessing an Elenan Ostionoid component, corresponds to the development of ceremonial architecture at the site. The importance of the site of Tibes is that before its discovery scholars were strongly debating whether ceremonial structures in Puerto Rico belonged exclusively to the

266 / L. Antonio Curet and Joshua M. Torres late pre-Hispanic period or whether their origins had an earlier antecedent. The site of Tibes not only produced ceramics and radiocarbon dates of the earlier period, but also it lacks evidence of a clear late occupation. In short, the structures seem to have been constructed before a.d. 1200 (González Colón 1984). Thus, Tibes demonstrated without any doubt that late pre-Hispanic ceremonial structures such as the ones in Caguana were preceded by similar structures in previous periods. Another important aspect of Tibes was pointed out by Oliver (1998), who noticed that the spatial arrangement of structures at Tibes is very similar to the one present in the later site of Caguana. From this observation, Oliver argues that both sites followed similar spatial canons in their construction and, because of the age of Tibes, these spatial symbolic canons had their origin in Pre-Taíno times. All this information suggests that major developments in symbolic structures and organization of society characterized by deeply symbolic structures and dynamic regional relationships occurred centuries prior to the development of the polities observed at the time of European contact. Since the discovery of Tibes, research has been conducted in other sites with ceremonial structures. As mentioned above, Oliver and Rivera Fontán have been conducting a variety of research in Caguana and its surrounding area. Oliver (1998, 2005) also studied the petroglyphs in the main plaza of this site and interpreted them by combining the archaeological and ethnohistoric data with ethnographic analogies from South America. Rivera Fontán (1992) also conducted a shovel-testing program in Caguana and was unable to detect significant trash deposits, which tends to support the ideas previously proposed by Alegría (1983:87) that these centers were in fact vacant. On another project, Rivera Fontán and Silva (1997, 2002) have been working for the past several years at the Batey del Delfín in Mayaguez, one of the few sites with ceremonial structures reported for western Puerto Rico. This structure is rich in rock art representing marine fauna and Rivera’s work has discovered several deposits and post molds potentially indicating the presence of domestic structures. Finally, Rodríguez Meléndez (2007) worked on one site with ball courts in the region of Jayuya (Mutaner) and one in Utuado (Sonadora) as part of her dissertation. Among other topics of interest, Rodríguez Meléndez attempted to study the activities that actually took place in the ceremonial structures. Arguing that archaeologists have traditionally relied on the historic descriptions, she conducted an excavation program around the structures in order to obtain archaeological evidence of the activities that took place in the structures.

Issues in the Study of Ceremonial Structures The brief summary of studies of ceremonial structures in Puerto Rico presented above shows many of the advancements in their interpretation. However, it also brings up some of the main issues that need to be addressed to gain a better under-

Plazas, Bateys, and Ceremonial Centers / 267 standing of their role among ancient social groups. Here we discuss four of these issues that warrant in-depth consideration. Vacant vs. Occupied Ceremonial Centers. The implications of the issue of vacant vs. occupied centers cannot be overstated. To begin with, it is important because the presence of domestic occupations associated with these sites entails a relational connection between social practices and the inherent structure and negotiation of power relationships in communal groups. These negotiations have recently been recognized as being central to understanding the processes of regional consolidation in the Caribbean (Torres 2005) and other parts of the New World (Pauketat 2000, 2007). A debate about these types of ceremonial centers has been going on in Puerto Rican archaeology since the early studies of Alegría (1983) in Caguana. The idea of a vacant center is not unique to the Caribbean and probably the bestknown case of a debate similar to this is that of the Maya centers. However, because the debates took place years ago, many Mayanists and Mesoamericanists consider this topic outdated and not an important issue in the study of ceremonial centers. This is unfortunate since the issue of vacancy vs. habitation is not trivial but, on the contrary, a critical aspect in the use of these sites that needs clarification in order to understand the relationality between domestic and ritual practices in the formation of communal identities. For example, an inhabited center has the strong possibility that it represents a seat of political power, economic control, or social and cultural focus, while a vacant center may indicate a neutral space for debate and negotiation and one where communal activities are conducted to increase cohesiveness and solidarity. In the case of the Puerto Rican ceremonial centers, this issue has not been completely explicated even though new information is available. In one case, work conducted by Rivera Fontán (1992) in Caguana was unsuccessful in discovering domestic deposits, which could indicate that the site did not have a permanent occupation. Congruently, with the information available, it is still unclear whether the early ceremonial center of Tibes was vacant or inhabited (see below). Two main issues requiring attention in addressing this problem are (1) the difficulty of distinguishing between household and ritual deposits such as feasting and ceremonial refuse and (2) the possibility that some centers, across space and/or time, may have been vacant while others were not. Time and Transformation. The first step in understanding ceremonial centers entails a critical examination of the formation processes associated with their development. To do this it is necessary to consider the rhythm and tempo of the physical construction of ball court and plaza structures, the use life of these sites, and the gamut of formation processes responsible for their current condition as part of the archaeological record. Understanding these factors provides us with the context within which to frame our interpretations. And although evidence is available for at least Caguana and Tibes indicating that their organization and

268 / L. Antonio Curet and Joshua M. Torres structure changed through time, in many cases these processes are not incorporated in our modeling or analytical explanations. In other words, the site as we see it today is often considered as a static plan, consistent in its present physical representation from its inception to its abandonment, and not as the result of a long historical process. Thus, while many recognize that this type of site has antecedents, ceremonial centers are treated as nonhistorical entities. This is problematic, as research of similar processes related to the formation and development of ceremonial spaces in other parts of the Americas has presented the potential connections deep in their spatiotemporal settings that influence the material and social outcomes much later in time (Sassaman 2005). Therefore, it is important to consider the historical processes that led to the formation and maintenance of the ceremonial centers in order to understand their social, economic, and political role among social groups. When considering aspects of chronology or trends through time archaeologists have tended to make use of the sociotemporal periods established by Irving Rouse. These periods are defined according to the timing of changes in one or more particular material indexes such as ceramics and lithics. Once periods are defined for a particular culture it is assumed that transformations in the rest of the cultural assemblage occur at the same rate. This is the case for ceremonial centers: they are assigned to a particular period and it is assumed that they were constructed at the beginning of the period and abandoned at its end. However, recent research by Rodríguez Ramos and colleagues (Rodríguez Ramos 2007; Rodríguez Ramos et al. 2010) has shown that the ceramics upon which the temporal frameworks are based are problematic and coarse in their resolution for assessing fine-scale chronological changes. Further, our conceptualization of time based on these essentialized frameworks has a tendency to homogenize our perception of the historical processes related to the development of these spaces that were constantly being transformed through their use and therefore constantly in a state of emergence. Regionality. Perhaps one of the most pressing issues related to our understanding of ceremonial centers is their role within the local and regional systems of which they were a functioning part (as discussed by Torres in this volume). As the word center implies, these types of sites are only a piece of a larger and regional network of social relationships. This entails a functional examination not only of archaeological sites and their geographical distributions through time and at varying spatial scales (i.e., local vs. regional) but also of the ways in which places are socially constructed and imbued with meaning. Therefore, from their inception (i.e., construction of structures) to their final abandonment, ceremonial centers are the focus of social relationships that connect multiple sites or settlements. When this is recognized, the question arises about the nature of this relationship between the center and the rest of the regional network. Do these multi-site net-

Plazas, Bateys, and Ceremonial Centers / 269 works represent at least ritual communities, in which several settlements or other social institutions share the ceremonial spaces and activities? Do they also represent social, political, or economic communities? So far, in Puerto Rico the assumption is that ceremonial centers are political, economic, social, and religious centers, even though few have tested this assumption. Tantalizing evidence suggests that they are social and religious centers, but no convincing data have been collected indicating that they are also political and economic centers, or seats of power. However, the lack of this evidence may be more a product of the scarcity of studies dedicated to this topic than of the absence of data. Finally, a third question is, what are the internal dynamics of ceremonial centers and how are they related to the immediate and regional social, economic, and political organization? Of course, the answer to this question is going to depend on whether the center is vacant or inhabited. If vacant, it has to be determined whether it was controlled by one of the settlements in the immediate region or its use was shared by several of them. If inhabited, then the internal social organization and relationships of the inhabitants have to be determined. Either way, the internal dynamics of the center influenced and at the same time were influenced by the intraregional interaction. Therefore, an understanding of the internal and external interaction of the centers is necessary to gain a better understanding of the social, political, and economic relationships of these groups. Unfortunately, in Puerto Rico most ceremonial centers have been treated within a single interpretational domain and very few studies have concentrated on studying the internal social organization of the site. In the same way, from the perspective of the region, ceremonial centers are seen as primate centers of systems that are monolithic in nature without ruling out that they may be systems composed of a number of different entities or factions. Scale. Finally, it is critical to understand the nature of these relationships and the roles of these sites at varying scales. Structurally, we assume that these physical spaces held similar functions at different levels of the regional settlement system in which performative ritual played an important role in the constitution and negotiation of power, identity, history, and place. However, the material representations of these structures in varying sizes and sizes of sites through time would seem to indicate (as mentioned previously in the example comparing Tibes and Caguana) a deep structure of meaning that lay the foundations for regional identity and social networks at varying scales in the form of performative or theatrical ritual (Geertz 1980). In the pages that follow we discuss the questions presented here using Tibes as our case study. However, these issues are complex and difficult to answer, and we do not claim that we can definitively answer any one of them in the case of Tibes. However, we can discuss them under the light shed by the data presented in this

270 / L. Antonio Curet and Joshua M. Torres volume. And, although the use of these data raises more questions than answers, it is a contribution to the debates in the discipline about these issues and to the understanding of Caribbean societies.

Vacant vs. Occupied Ceremonial Centers As mentioned above, the natures of occupied vs. unoccupied ceremonial centers reflect major differences in social and political organization. In this section we want to present some of the methodological and empirical problems involved in answering this question and discuss them within the context of Tibes and the data obtained to date. In order to answer this question, archaeologists have to be able to distinguish between the two types of ceremonial centers in the archaeological record and must address the implications of shifting site functions through time. Particularly, we need to focus or concentrate on identifying the different activities associated with each one of the types. Because of the nature of the activities it is clear that both types of centers share some of these activities, especially those related to communal events including feasting and ritual ceremonies. Therefore, probably the clearest type of activity that can be used to distinguish one type from another is the permanent domestic activities in occupied centers, contrasting with the temporary and limited domestic activities in the vacant ones. While this sounds pretty clear and straightforward, in practice things are not that easy, since remains of repetitive feasting, rituals, and temporary residences can, at least at first, look very similar to permanent domestic units archaeologically. In our opinion, there are three different types of activities that need to be distinguished in order to differentiate between vacant and occupied centers, with specific focus on the material evidence for different domains of social practice related to (1) feasting, (2) ritual, and (3) domestic activities. In general, there are three pieces of evidence applicable to archaeologically distinguishing among the three: architecture (i.e., buildings), artifacts, and food remains. In terms of architecture, one would expect occupied centers to have permanent domestic structures, as opposed to more provisional ones at vacant centers. However, if vacant centers are used one or more times every year, it is possible that there would be more investment in the structures to avoid constructing new buildings in every visit and, instead, conduct repairs regularly. This possibility makes discerning between temporary and more permanent occupations difficult at best. Further, living structures in vacant centers may have different forms and sizes than the average domestic building in habitation sites. For instance, vacant centers may have larger living structures to accommodate the large number of participants involved in the ceremonies. Also, the living areas of this type of center may not include all the activity areas that accompany normal, permanent household clusters. Of

Plazas, Bateys, and Ceremonial Centers / 271 course, this is dependent on the life span of the event, since longer stays may actually produce a similar number and type of activity areas compared to those present in permanent households in order to fulfill the necessities of the groups during their extended stay at the center. Because most buildings in the Caribbean were built with perishable materials, it is difficult to discover and study them in any kind of site. Most of the time these buildings are detected and reconstructed based on postholes and, in some cases, by the presence of hard-dirt floors. For this reason, only a few domestic structures have been studied in the region, making the possibility of comparing structures from ceremonial centers to the ones in habitation sites impractical at this point. In Caguana, Mason (1941) identified large numbers of post molds, and he and later Alegría (1983) found actual posts that preserved well in the clayey humid conditions of the soils. While it was difficult to discern the shape of most of the structures, a large round building was discovered between the main plaza and the largest ball court. However, it is difficult to discern whether any of the post molds, including those of the round structure, actually belong to domestic units or to specialized ceremonial buildings. In Tibes, several post molds and at least one carbonized piece of a post have been found by the excavations of both the Sociedad Guaynía and the present project. All of them seem to belong to the late Elenan Ostionoid period (i.e., after a.d. 900). However, so far excavations have not been able to determine the size, shape, or nature of the structures. Although some of them seem to be associated with food remains and tools, it cannot be determined whether they belong actually to permanent household structures. Furthermore, it has not been established whether these post molds belong to buildings older than or contemporaneous with the ceremonial structures, that is, whether they were being used when the site was already a ceremonial center. Another piece of evidence that can be used to distinguish between the two types of centers is the artifacts. In general, one would expect major quantitative and qualitative differences in artifact assemblages: vacant centers should have lower densities of utilitarian materials, a narrower range of artifact types, and, possibly, more “specialized” types of artifacts used in ceremonies, rituals, and feasts and not normally found in habitation sites (e.g., large cooking and serving vessels and ritual objects). In the case of occupied centers, we should expect the whole range and high densities of domestic artifacts plus the ones related to ritual and feasting ceremonies. In the case of Tibes, it is difficult to make comparisons of this type of archaeological evidence because not all the appropriate analysis necessary to compare the artifact assemblages is completed and, in part, because detailed information is not available from other habitation sites in the region. Although a thorough analysis has been conducted on the ceramics, the study was directed more to understand the chronology and cultural affiliation of the deposits, to characterize the cultural

272 / L. Antonio Curet and Joshua M. Torres context of the different pottery assemblages at the site, and to determine the nature (e.g., domestic, ceremonial, production) of the deposits. Despite this lack of evidence, it can be said that Tibes does not seem to have a high amount or density of decorated pottery. At least, the pottery does not seem to vary significantly from assemblages found in habitation sites. Further analysis on the ceramic collections will concentrate on aspects of specialized forms and size of vessels, compare the decoration found in Tibes with that in other collections in the region, and determine changes in pottery assemblages through time, but within each period. Similarly, Walker’s analysis of stone artifacts (this volume) has concluded that in terms of lithic assemblages, those collected from Tibes represent neither a large quantity of lithic items nor a diversity of materials that would indicate the variety of activities suggestive of long-term domestic use or of specialized activities. Lithic assemblages at Tibes, in general, seem to be as common as those of any habitation site in the island. Even items of religious paraphernalia such as threepointers do not seem to be more prevalent, bigger, or more elaborate than in other locations. Perhaps the only significant difference in stone artifacts is the presence of “sacrificed” three-pointers (three-pointers with one point missing). Our impression is that this “commonness” is also true for utilitarian and religious artifacts made of shell or mother-of-pearl. Although they are present in Tibes, they do not seem to be exceptionally special. Finally, bone artifacts found by this project are very few in number and most of them are ornamental (e.g., shark teeth used as pendants). Differences between habitation sites and vacant centers should also be discerned through significant variability in the faunal and botanical assemblages. Because occupied ceremonial centers include remains from both domestic and ceremonial foods, we should expect them to show higher densities and diversity of faunal and botanical remains. In the case of vacant centers we should expect the opposite. However, because they occur at special occasions, ceremonial meals (ritual or feasting) tend to follow canons that can include the use of special resources or preparation processes. This is especially true in the case of stratified societies, in which the elite may consume different food than the rest of the population. Therefore, in the case of some exceptional food, it is expected that ceremonial loci will contain more limited, special types of foods in higher densities (i.e., specialized food used in ceremonies and feasting). Interestingly, both the faunal (see deFrance et al., this volume) and botanical (see Newsom, this volume) records of Tibes show a wider diversity of species than other sites in Puerto Rico, even those around Tibes. This seems to match the expectation that Tibes was an occupied center at least during some of its history. Newsom and deFrance et al. also relate this high diversity to the presence of a well-developed elite stratum and of ceremonies that required special types of foods. However, both studies warn that these patterns may have been produced by

Plazas, Bateys, and Ceremonial Centers / 273 the thorough research design of the Tibes project that, contrary to most projects on the island, paid special attention to the recovery of food remains (see discussion of botanical analysis below). In the case of faunal remains (see deFrance et al., this volume), the great majority of the food deposits at Tibes seem very similar to domestic deposits of other sites. However, some differences or special distributions can be found within the site. For example, it is interesting that four exotic species of small terrestrial mammals were found at the site. Three of them were present in the early Saladoid period, while two were found in the later Elenan Ostionoid period. Spiny rat (cf. Heteropsomys sp.), rice rat (Oryzomys sp.), and hutia (Isolobodon portoricensis) were discovered in Saladoid deposits in the north part of the site. The first two species consist of one MNI each, while the third is more prevalent. This is the first time that spiny and rice rats have been reported for a site in Puerto Rico, although they are not uncommon in the Lesser Antilles (Newsom and Wing 2004). Hutias and guinea pigs were discovered in the Elenan Ostionoid deposits. While hutias tend to be most abundant in these deposits, guinea pigs are not as uncommon as the other rodents, but they still occur in significantly lower numbers. Hutias were brought from Hispaniola, probably in Saladoid times. Their abundance in the deposits of all periods at Tibes and, as a matter of fact, in the whole island, strongly suggests that they were used as a common source of food in many contexts. The species is so prevalent that Newsom and Wing (2004) have suggested that they may have been tended, possibly in domestic contexts. Therefore, hutia by itself cannot be considered a special type of food consumed on special occasions, unless it was prepared following a different process than in daily meals. Spiny rat and rice rat could have been imported from the Lesser Antilles or the continent, while guinea pigs were probably brought from the latter as they are absent in the former. The presence of these exotic fauna since Saladoid times suggests that ritual or feasting activities may have been taking place at the site since that time. It is not uncommon for nonstratified communities to conduct ritual ceremonies and feasting that may include special types of foods such as these species found at Tibes. Within these contexts, special food can be used in feasting and ceremonies to increase communal integration and, at the same time, to differentiate between groups by facilitating competition and aggrandizement (e.g., differential access to exotic food). If it is true that processes such as these were active since Saladoid times, then it is possible that the development of a ceremonial center at Tibes is not a coincidence but, on the contrary, part of a trend that began early in its existence. Hutias seem to continue as a source of protein during the Elenan Ostionoid period, but they are more prevalent than before, a trend that is common in other areas of Puerto Rico (Miguel Rodríguez, personal communication 2005). However, it is during the early part of this period (i.e., Monserrate times, around a.d. 900) that

274 / L. Antonio Curet and Joshua M. Torres guinea pigs appear in the archaeological record of Tibes. Guinea pigs were found for the first time in Puerto Rico on the north-central coast of the island (Wing 1996; see also Quitmyer and Kozuch 1996) and since then they have been reported for other sites. However, in all cases their MNIs tend to be relatively low. It is significant that these animals appear in the archaeological record in small numbers despite their ability to reproduce in great numbers. This suggests that their reproduction and access to them were highly controlled by someone. However, with the information available on their distribution and their place of origin it is difficult to determine whether they were being reared in Tibes, the region around it, Puerto Rico, the Caribbean islands, or the continent. Independent of where they originated, their presence in a ceremonial center like Tibes strongly indicates that they were used in either ritual activities or feasting, or in both. Another aspect of faunal evidence that may be related to feasting is the discovery of large pockets of Turritella variegata shells in a cooking midden (OP19 units; see Curet, this volume). The high concentrations in a location near the main plaza suggest that this species may have been preferred as a foodstuff in ceremonies or special meals. However, a more thorough comparison has to be made with other sites in the region and the island to verify this trend. The botanical remains can also shed some light on differences between Tibes and habitation sites. The recovery and analysis of macrobotanical remains at Tibes have been extremely successful. While few botanical foodstuffs per se have been found in Tibes, a large amount of carbonized pieces of wood was recovered, some of these belonging to fruit trees. Surprisingly, very few seeds have been found and not all of them seem to belong to edible species. For these reasons, most of the discussion that follows focuses on wood identification. As mentioned before, one of the most interesting results of the macrobotanical analysis is that wood variability at Tibes is one of the greatest so far obtained in any Caribbean site. This is difficult to interpret. In part, it is easy to assume that this must be related to the ceremonial nature of the site, its long history, or the possibility that it was the seat of a powerful polity. However, there is no reason to believe that any of these assumptions are necessarily true. If Tibes was primarily a ceremonial center one would expect a narrower range of species due to its specialized function. In other words, few types of activities will have taken place at the site, requiring a limited number of different species of wood. It is not clear why either a site with a longer history or one that is the seat of a polity needs to have a broader number of species of wood trees than other sites, unless some of the activities changed through time or took place mostly in political centers. Even then, however, the remains of these activities should be overshadowed by normal, and more conspicuous, household activities. Another possibility is that cultural preferences for different woods changed through time from the early Saladoid to the late Elenan Ostionoid. However, this high variability in wood species is not found

Plazas, Bateys, and Ceremonial Centers / 275 in other sites with sequences as long as or longer than Tibes. Moreover, the wood variability at Tibes may be a reflection of the high variability of botanical species in the southern dry forest of Puerto Rico, but again, this diversity is not found in other sites in the region. Alternatively, the large number of species may have been produced by a combination of possibilities. In the case of Tibes, it is possible that it was an occupied ceremonial/political center, where different woods used in both domestic and ritual contexts were present. A final possibility is related to the recovery strategy used at Tibes compared to that of other projects. As mentioned in Chapter 3, the research design of the project was developed in consultation with the specialists from its beginning. Special attention was given to the recovery process and guidelines. In the field the strategy was altered according to the field conditions and the nature of the discoveries, always to maximize recovery and increase the preservation of the fragile botanical remains. Very few projects in the Caribbean have followed such a thorough process. Thus, it is difficult to determine whether the differences in the number of wood species identified between Tibes and other sites in the region are real or a product of differences in the intensity of the recovery process. In addition to species variability, the paleobotanical analysis has provided interesting insights about botanical resources used at Tibes. While very few seeds have been discovered at the site, the evidence obtained from charred wood has provided a great body of information on firewood and fruit trees. However, the discussion here concentrates on two species that are strongly related to ceremonies and ritual activities: cohoba (Anadenanthera sp.) and evening primrose (Oenothera sp.). An interesting aspect of these two taxa of plants is that they are not native to Puerto Rico or the Greater Antilles, but they seem to have been imported and tended in home gardens. According to the early European chronicles, cohoba seeds were the main hallucinogenic ingredient used in ceremonies to contact the supernatural. While the use of cohoba in earlier times has been always inferred from the presence of the paraphernalia used in these ceremonies (e.g., vomic spatulae, inhalers), this project is the first to find direct evidence of the tree at a site. To a certain degree, the presence of cohoba in a ceremonial center such as Tibes is not surprising. However, what is surprising is its pervasive presence in most of the deposits at the site, indicating that this tree was being used since the foundation of the settlement in Saladoid times. The second species, evening primrose, is a mild narcotic and medicinal plant. Newsom (1993; Newsom and Wing 2004) has detected its presence in many sites across the Caribbean, even though its modern distribution is highly restricted to a few regions. Contrary to cohoba, this taxa is present in low numbers and almost exclusively in Saladoid deposits. Therefore, this evidence suggests that this plant was not being used during the ceremonial center “phase” of the site. Nevertheless, in

276 / L. Antonio Curet and Joshua M. Torres conjunction with the presence of cohoba, it also indicates that ceremonies that involved hallucinogenic drugs were conducted at the site since early times. When considered together, the botanical and faunal data provide mixed signals in terms of the nature of the ceremonial center. The extremely low number of seeds seems to suggest that a great deal of the food processing was being conducted outside Tibes, probably in other sites. This supports the possibility that the center was a vacant center. However, the widespread presence of charred wood from fruit trees seems to contradict this. The “commonness” of the faunal assemblage, the broad diversity of botanical and faunal species, and the presence of ritual-related resources suggest that the site was both a habitation and ceremonial location. Summarizing, the multiple lines of evidence recovered from Tibes point in different directions in our attempt to determine whether the site represents a vacant or occupied ceremonial center. Some points of evidence, like that coming from artifact assemblages, do not show any special qualities that characterize a vacant center. On the contrary, all around with few small exceptions they look as mundane as those from any other site. The faunal and botanical remains also show some evidence of habitual life, but, at the same time, they provide evidence of feasting or ritual food. While this latter evidence has been found near ceremonial structures, it is still difficult to determine whether the deposits represent ritual or ceremonial assemblages or a variety of activities including domestic ones. One issue that needs to be dealt with in order to solve the issue of vacancy or occupation is the time dimension. It is important to, first, discern which deposits and activities are contemporaneous or not and, second, determine whether the nature of the activities, the deposits, and the site as a whole changed through time. In other words, the nature of the site and ceremonial center may have shifted throughout its history. These aspects of the archaeological record of Tibes are discussed in the next section.

Time and Transformation Ceremonial centers are not static entities that are created all at once and remain unchanged until their abandonment. They are dynamic places that shape and are shaped at the same time by the social practices associated with them. They are liminal spaces that are materially constructed and socially constituted. They can be interpreted and reinterpreted according to the situational contexts of their use by social agents in the process of negotiating their social realities. The creation of ceremonial centers is therefore not the result of a single event. Rather, they are the result of historical and dialectical processes that interact in multiple social dimensions. For example, reasons for the conception of a ceremonial center in a particular place may be related to one or more factors such as being one of the oldest ancestral settlements, a traditional sacred location, the site of a particular “mi-

Plazas, Bateys, and Ceremonial Centers / 277 raculous” event within the historical context of the group, or the seat of a powerful polity competing for ideological and symbolic capital (Siegel 1999). Why a center is founded in a particular place is difficult to determine. However, it undoubtedly involves the intersections of history, social memory, a system of beliefs, place, and the material worlds within which social actors dwell. Despite the difficulties of measuring the intangibles of ceremonial centers, archaeologically we can detect trends and evidence of the social and historical factors that were involved in their foundation and that can provide clues to better understand their formation, use, function, and role in the social history of the group. In order to accomplish this, however, we have to control for several factors. Specifically, we want to discuss cultural processes and aspects of chronology, all of which are critical in order to reconstruct the historical trends in an accurate and precise manner. Cultural formation processes are factors produced by human groups that created the archaeological record as we see it today. As mentioned in Chapter 2, these factors include the social, cultural, economic, and religious activities conducted by the original inhabitants of the site, any changes produced by these same inhabitants, and changes produced by other people after the abandonment of the site or during reoccupation. Here we emphasize cultural processes active during the occupation of the site (i.e., systemic context). Finally, several aspects of the way we measure time need to be addressed as well. While many 14C dates are obtained from archaeological sites, the great majority of dating in archaeological projects relies on association with known dated materials such as ceramic types or styles. However, this procedure tends to assign archaeological features or events to broad periods. This approach has three consequences. The first is that some events and archaeological finds can be assigned to one part of the archaeological period and not to its whole span. Second, this practice assumes that cultural (e.g., ceramic styles), social, political, and economic changes occurred at the same rate and at the same time. Third, in the case of the Caribbean, we rely on periods defined by Rouse (1986, 1992) that have been questioned recently by comparisons of radiocarbon dates with ceramic types and styles (Rodríguez Ramos 2007, 2008; Rodríguez Ramos et al. 2010). In other words, recent studies have shown that some of Rouse’s periods may have lasted longer and overlapped with later periods, suggesting the coexistence of cultural traditions that normally were considered to belong to different times. However, in Tibes we do not have yet the information (e.g., a large number of radiocarbon dates from a large number of deposits and levels) necessary to deal with this last issue. Suffice it to say, for now, that we are aware of the problem and will use Rouse’s categories for the time being to refer to the cultural traditions that existed during some periods. We do not assume that the traditions are sequential and agree that some of these traditions may have overlapped. As much as the data

278 / L. Antonio Curet and Joshua M. Torres allow us, we make use of radiocarbon dates that we have available to contextualize materials instead of relying on the general periods developed by Rouse. In addition to sociohistorical factors, we need to consider the behaviors of social agents when trying to reconstruct the social history of Tibes. Tibes seems to have been inhabited during Saladoid times, but, unfortunately, we lack appropriate radiocarbon dates to suggest a beginning date for the site. Although some Hacienda Grande pottery is present, the prevalent component of the series is Cuevas material. Traditionally, this latter ceramic style dates between a.d. 400 and a.d. 600. However, the studies of Rodríguez Ramos (2007, 2008) have shown that it may have lasted several centuries longer. Because of the presence of two burial clusters during this time and the lack of strong evidence for social stratification, this group has been interpreted as relatively egalitarian. Tantalizing evidence for ritual or high-status food is present in the form of exotic animals such as spiny and rice rats. As mentioned before, the presence of these species may be an indication that the level of ceremonialism among Saladoid groups may have been more elaborate than many have argued (see also Siegel 1992, 1996). If so, this ceremonialism may have been part of the origin of the processes that led to the conversion of Tibes into a ceremonial center. The prevalence of Cuevas over Hacienda Grande style ceramics suggests a stronger occupation during the late Saladoid than during the early part of it. Evidence for Hacienda Grande deposits is based solely on the presence of ceramics. No study has been conducted in the early deposits to determine that a true occupation was present. Consequently, to date we cannot conclude definitely that Tibes had a Hacienda Grande occupation, and it is equally possible that the site was only visited sporadically or for short periods during this time. Interestingly, most of the ceramics identified as Hacienda Grande belong to very elaborate and decorated vessels, including effigy vessels and pedestal bowls, possibly used for ritual activities. This suggests that if the location of Tibes was not a habitation site during the early Saladoid period, then the visits and short stays included ritual practices of some kind. The distribution of the ceramics indicates that most of the Saladoid deposits are located in the northern half of the site. The only exceptions to this are the clusters of burials reported by Alvarado Zayas (1981) and González Colón (1984), which are located in the center and southwest portions of the site. Because of the presence of these burials and the high density of material culture, it has been assumed that Tibes was a habitation site during this period. However, no evidence for domestic occupations for this period has been discovered so far. One possible reason for this is that most of the excavations conducted by the original work of Sociedad Guaynía and by this project have concentrated in the central part of the site around the main plaza, which is dominated by Elenan Ostionoid assemblages.

Plazas, Bateys, and Ceremonial Centers / 279 Future excavations are planned to expand investigations to the northern part of the site to better define the Saladoid occupation. An increase of the site area, especially in the central and southern parts of the site, is observed during Elenan Ostionoid times. It is believed that some time during this period the ball courts and plazas were built and the ceremonial center developed. The relatively rapid changes and the intensity of some of the activities that took place during this period make the archaeological assemblages difficult to interpret. These activities most probably included the destruction or movement of earlier deposits, the removal of old buildings and construction of new ones, and the movement of dirt and rocks especially from the leveling process during the construction of the ceremonial structures. Interpretation is also complicated by our limitations in dating some of these events, especially the construction of the ceremonial structures, since they were kept relatively clean and very few artifacts or even charcoal are found inside them. Besides the presence of ceramics and food remains we know very little about the early part of the Elenan Ostionoid period (i.e., Monserrate style) at Tibes. Because of the cultural transformational processes and the reorganization of space late in time, it is unclear how space was organized and occupied during the early part of the period. It is also difficult to determine whether ball courts or plazas were built during this time, as suggested by González Colón (1984) and Siegel (1996, 1999). This last task is complicated by the possibility that early ceremonial structures were dismantled to build later ones, although it is equally possible that some of the structures may have survived until the abandonment of the site. The lack of information on the early Ostionoid period, however, could also have been produced by sampling issues, since most of the excavation units have been located on late Elenan Ostionoid deposits. Evidence gathered to date for the late Elenan Ostionoid indicates that major cultural formation processes took place during the development of the ceremonial center that impacted earlier assemblages. This evidence includes the presence of secondary, mixed deposits next to several of the ceremonial structures, postholes containing trash, and finds suggesting that some parts of the site were leveled with material from old trash middens. Because of the convergence of the radiocarbon dates obtained from these pieces of evidence, we have suggested (Curet et al. 2006; Curet et al. 2003) that a major reorganization of the site occurred sometime between a.d. 1000 and a.d. 1100 and that this probably involved the formalization of the plan of the site as we see it today. Thus, while most probably the processes were operating much earlier at Tibes, it is not until late in the sequence that Tibes went through a transformation in terms of the physical organization of space leading to its development as a ceremonial center. Interestingly, chronologically, these major changes do not neatly coincide temporally with the periods defined by

280 / L. Antonio Curet and Joshua M. Torres Rouse (1952, 1986, 1992) for the Elenan Ostionoid (i.e., a.d. 900–1200), but they date to the middle of the period. This demonstrates that ideological, political, cultural, social, and economic changes are not even, linear processes but rather a product of transformations in social practices that are cumulative and historically contingent, as argued by Bourdieu (1977) and more recently by Pauketat (2000). As mentioned above, the distribution of the ceremonial structures at Tibes is very similar to the one found at Caguana, indicating that their arrangement has some meaning and significance and that in both cases they followed similar ideological canons. There are several interesting observations about the location of some of the structures. The first one, which has been pointed out on many occasions by many of us (Alvarado Zayas 1981; Curet 1992a, 1996; Curet et al. 2006; Curet and Oliver 1998; González Colón 1984), is that the main plaza and the largest ball court (Structures 6 and 3, respectively, in Figure 1.3) are located over a Saladoid (Cuevas) cluster of burials. Curet and Oliver (1998) argue that this is not a coincidence and that it is probably evidence of ritual continuity from the Saladoid to the Ostionoid traditions and the institutionalization of the ancestor cult during the latter period. A second observation is that while most of the Ostionan deposits are located in the central and southern parts of the site, the northernmost ball court (Structure 9 in Figure 1.3) is located over purely Saladoid deposits. While we are waiting for radiocarbon dates from this area, the ceramics tend to suggest that this area may include the oldest deposits of the site. It is, also, in this area where some of the Saladoid ceremonial pottery was found by the Sociedad Guaynía. A final observation on the distribution of the ceremonial structures is that two small ball courts and the southern part of the largest one are located in an old river channel (Structures 1, 4, and 3 in Figure 1.3). This channel has not been dated, yet, but it obviously became inactive before the late Elenan Ostionoid. But, considering that most of the early deposits are located on the northernmost part of the site and later deposits tend to shift further south through time, it is possible that this was the location of the Portugués River in Saladoid times. This may explain the spatial distribution of deposits belonging to different periods and why it is possible for the ball courts to be located there. A geomorphological study of the site, including this channel, is planned for future field seasons. In addition to the ceremonial structures, we also have evidence for the Elenan Ostionoid period of at least one building, in the form of postholes and a hardened dirt floor, and a cooking area, based on high concentrations of food remains (faunal remains), some of them burnt, ashes, fire-cracked rocks, charcoal, large pieces of griddles, and a burnt occupation surface. González Colón (1984) also reports the discovery of postholes for this period in other parts of the site. Initially, in our case, it was believed that both the building and the cooking area were part of domestic or household units, supporting the idea of an occupied ceremonial center.

Plazas, Bateys, and Ceremonial Centers / 281 However, considering their proximity to some of the ceremonial structures and some indications of feasting, it is possible that these archaeological features may have been related instead to the ceremonial activities. Tibes seems to have been abandoned at the end of the Elenan Ostionoid period. The most recent date obtained from the site is a.d. 1270 (calibrated, see Curet, this volume). This date was obtained from the top layer of a large pit discovered near the northwest entrance of the main plaza. The pit did not contain a high density of artifacts and its function is not clear. However, several ceramics with traditional Chican Ostionoid or late Ostionan Ostionoid designs were obtained from throughout the pit. We are expecting the results of more radiocarbon dates from the bottom of the pit to test whether it represents a single event. Nevertheless, the lack of a clear deposit belonging to the Chican Ostionoid period strongly suggests that the site fell into disuse shortly after the end of the Elenan Ostionoid period, probably around a.d. 1200–1300. The reasons for the abandonment of the site are still elusive. González Colón (1984) believes that the thick layer of river silt that covers the cultural strata over some parts of the site is evidence of a large flood that forced people to abandon the site. However, this layer is located only on the western part of Tibes, indicating that not all sections of the site were flooded. Furthermore, soil studies by Scudder (2001) have indicated that multiple floods similar to the ones observed occasionally at the site today may have produced this layer. Thus, no abnormally large flood was responsible for depositing the alluvial soils. Finally, the abandonment of Tibes is not unique and localized. Many sites over the south coast of Puerto Rico, especially sites with ceremonial structures in the foothills, were abandoned more or less at the same time as Tibes (Lundberg 1985; Rodríguez 1985; Torres 2001). However, more recent research conducted by Torres (2008 and in this volume) and recent work at Jacana north of Tibes (Espenshade et al. 2007; Espenshade and Young 2008) indicate that several other settlements on the Portugués River were occupied at least into the early Chican Ostionoid. This suggests that whatever explanations are offered, they have to take into consideration shifts in populations on the southern coast (Curet 1992a, 2005; Curet et al. 2004). Summarizing, Tibes has a long history that culminated in the formation of the ceremonial center that we know today. However, its form, function, and role changed through time. During Saladoid times, we have been assuming, the site was a village, but to date no conclusive data have been obtained to support this argument. It is possible that Tibes was the location for visits that may have included some rituals and, at least toward the end of this period, group communal activities. Either way, based on ritual ceramics and the presence of some exotic fauna such as spiny and rice rats, it seems that ceremonies and rituals were already taking place at this location prior to the end of the period. Later on we see the cultural

282 / L. Antonio Curet and Joshua M. Torres changes that led to the development of the Elenan Ostionoid tradition, but this is mostly based on changes in ceramic styles. During the early Elenan Ostionoid (i.e., Monserrate style) in Tibes, so far the only other change we can correlate with this trend in ceramics is a shift of deposits moving further south toward the central part of the site. It is unclear whether Tibes was or was not a habitation site at this time, or whether people were still being buried at the site. It is possible that some ceremonial structures were built at this time, but there is no definite evidence for this. Ceramic changes and further shifts of deposits to the south also characterize the late Elenan Ostionoid period (Santa Elena style, a.d. 900–1200). We hypothesize, however, that the ceremonial center did not develop until much later, possibly between a.d. 1000 and a.d. 1100, and was relatively short-lived since it seems to have been abandoned between a.d. 1200 and a.d. 1300.

Regionality Regional-scale interactions among social groups are recognized as being fundamental to human sociality (Sassaman and Randall 2007). However, up until recently, most ceremonial centers on Puerto Rico were treated as isolated features on the landscape that could be studied and explained based solely on their own assemblages. Even when Alegría (1983) argued that ceremonial centers such as Caguana were vacant centers, most scholars ignored the surrounding support population in their discussion. It is clear from Torres’s (2001, 2005, this volume) studies of southern Puerto Rico and Oliver and Rivera Fontán’s work around Caguana (Oliver 1998; Oliver et al. 1999; Rivera Fontán 1992; Rivera Fontán and Oliver 2005) that ceremonial centers are regional phenomena and they should be treated as such in our modeling of past human behavior. From a regional or subregional perspective ceremonial centers represented places of activity for surrounding settlements in the formation of communal bonds and the negotiation of identity and power relationships. Settlement pattern analysis of the distribution of ceremonial centers is an initial step in the process of understanding these regional relationships. However, it is critical that we begin to move beyond descriptions of site distributions and take these studies further. First, studies focused at the regional level can provide the history of a landscape in which to form a framework for examining the temporal and spatial contingencies that dictate the interaction of social groups and the formation of ceremonial centers. Places have history and these histories are emergent from the way in which people construct their social realities within the spaces they inhabit (Gray 1999; Hirsch 1995; Ingold 1993). As mentioned in the previous section (Time and Transformation) it is necessary to understand these emergent histories and it is our contention that patterns of change in organization at the regional level will have correlates at

Plazas, Bateys, and Ceremonial Centers / 283 ceremonial centers that are instrumental in understanding their development, use, and importance through time. Second, an examination of the material culture from the settlements surrounding these ceremonial centers should yield clues to the differences in social identities and status through the variability in their function and in the material culture exhibited. Further, comparison of assemblages between the ceremonial centers and their surrounding settlements should yield clues to how groups were interacting within these ceremonial centers based on the archaeological contexts of the deposits. Archaeologists since Alegría (1983) have noted that ceremonial centers were for the meeting of multiple settlements for ritual activities. It seems only logical that we actually start substantiating these relationships archaeologically. Moreover, and by the same token, this comparative examination would entail smaller sites with ceremonial structures (e.g., El Bronce, PO-39) and should yield clues into the nature of the function of plaza/ball court sites that reflect the relationality of social practices at different scales. Finally, another aspect of critical importance is characterizing the spatial and temporal organization of ceremonial features throughout a given region. Regional studies of ceremonial centers can help elucidate some of these relationships, as spatiality is never random and in the case of competing social groups, patterns are produced within and between acting polities or communities (Smith 2003:75). Recent research has attempted to correlate the rhythm and tempo in the construction of ceremonial sites on the island with increases in social competition and increasing social complexity (Siegel 1999). Problematically, this perspective (as mentioned previously) assumes similarities in the function of ceremonial structures at the regional level through emphasizing presumed hierarchical power structures that may or may not be real. And even though the role of competition has been rightfully emphasized as a mover in the construction of these spaces and the politicization of communities in general, it does not address their function within the regional network of sociopolitical relations—not all ceremonial features on a landscape should be the same in terms of the social practices associated with them and their function at different scales. Rather, regional ceremonial structures may have performed different functions at different scales (and at different times) for integrating groups within a village, within multivillage communities, and within regions. The development of formalized ceremonial architecture along the south-central coast between the seventh and eleventh centuries, as is evident at the sites of Tibes, Las Flores, and El Bronce, certainly indicates a regional acceptance of similar religious practices. The development of this supraregional religious affiliation and its materialization in the form of plazas and ball courts suggest a larger transformation in social practices on the regional level.

284 / L. Antonio Curet and Joshua M. Torres Adler and Wilshusen (1990) have shown a clear relationship between size of ceremonial space and population. Clearly, if we assume that ceremonial structure size is related to group size and scale (i.e., local community vs. regional community groups) this aspect of scalar relations becomes clearer. However, existing assumptions that ceremonial centers represent the presence of chiefdoms ipso facto should be critically evaluated. On the island of Puerto Rico ceremonial architecture is relatively small and current research is suggesting that these structures could have been built without substantial labor investment. Even in the case of Tibes, one of the largest ceremonial centers on the island, critical examination of labor involved in the construction of these features will probably reveal that the human costs involved were quite small, minimizing their importance as a proxy for chiefdoms (Torres et al. 2007). In this sense, some of the patterns observed for the distribution of ceremonial architecture may actually have less to do with actual boundaries of sociopolitical units or chiefdoms and more to do with communal use of space as centers for surrounding settlements. The regional connections of these ceremonial sites are further brought to bear when we begin to examine the burials in the ceremonial centers at which human remains have been recovered. Recent research has discussed the potentiality for these burials coming from different communities surrounding the ceremonial center (Keegan 2009). Future research to critically evaluate this hypothesis is being conducted by Jason Laffoon using strontium isotope ratios (Laffoon et al. 2006).

Scale The issue of scale in analyzing Tibes is a difficult one to address at this point. In part this difficulty is in the nature of multiscalar analysis, wherein data at different levels and scales (e.g., from burials and households to regions, macroregions, and the whole island; from artifact and faunal analysis to the analysis of assemblages and settlement patterns) have to be collected in a consistent manner in order to make comparisons and interpretations. It is difficult in our case, because the detailed data necessary to make these comparisons are not available for many sites in the region, or most of the island for that matter. Nevertheless, here we present some relevant points that show the importance and necessity of this type of analysis in order to gain a fuller understanding of Tibes within the broader social and cultural phenomenon from which it emerged and of which it was a part. It is clear from the discussion above that Tibes operated at multiple scales. Each one of these had its own dynamics, they are not necessarily mutually exclusive, and they were influenced by and influential on the other scales and levels. These multivalent scales, materially evident in plazas (Heckenberger 2005:261), and the chaotic nature of population movement and social interaction through time and space in the Caribbean (Keegan 2004) emphasize the nonlinearity and multidimensional

Plazas, Bateys, and Ceremonial Centers / 285 nature of social organization and development. We see this, for example, in the relationship between Tibes and the surrounding area, where the recursive nature of the local and the regional was instrumental in shaping the type of center it came to be, while at the same time we can see that the social practices conducted at the site have material correlates to the broader region. In the case of Tibes, we have begun to understand the complexity of the assemblages and the multiplicity of processes that created them. At a larger scale, we are beginning to address these relationships by framing our inquiry into the position and social role of Tibes beyond its immediate region, including the rest of the Portugués River basin (including the enormous and long-lived site of Caracoles) to neighboring river basins, to the island, and beyond (Curet 2003; Rodríguez Ramos 2007). The implications of this multiscalar approach are critical for evaluating and constructing historical interpretations not only for the island of Puerto Rico but also for the prehistoric circum-Caribbean as well. The case of the whole island should not be underestimated. As mentioned before in this volume, Tibes is in the boundary region where the Ostionan and Elenan Ostionoid traditions overlap. Even though Tibes belongs mostly to the Elenan Ostionoid tradition, it is not uncommon to find Ostionan Ostionoid series ceramics at the site. This mix of materials wherein sites belonging to one cultural tradition always produce small numbers of sherds from another contemporaneous tradition is very common across the island. This pattern has been recognized by a number of researchers, but few, if any, have attempted to explain it. However, as Torres (this volume) argues, it is of vital importance to understand and explain this phenomenon in order to understand Tibes’s role at varying scales and the ways in which the peoples of the past constituted themselves socially within their world. Thus, it is clear that a social and cultural phenomenon such as Tibes cannot be considered to have operated in a social and cultural vacuum. Its influence is represented in its materialized association with a broader supraregional socioreligious complex of symbolism, meaning, and practice. Research regarding the construction of social realities in other parts of the world has shown the scalar relationships between symbolic constructions of space and performative action as part of the natural ordering of societies (Geertz 1980; Heckenberger 2005; Inomata and Coben 2006). Without a doubt, the relational nature of symbolic space and social practice, and transformations of them through time, at multiple scales, influenced the rise, development, and, possibly, abandonment of Tibes as a ceremonial center.

Conclusions So, we can say that Tibes falls into our definition of a ceremonial center and one used within the contexts of the prehistoric Greater Antilles in general. However,

286 / L. Antonio Curet and Joshua M. Torres what was its importance in the past? It is probably safe to assume that the site was part of a broader network of social relationships entailing the negotiation of power and identity for local community groups. This was materially represented in the formal delineation of ball courts and plazas and the performative practices associated within them. Importantly, Tibes did not physically emerge as it presently exists but is the product of hundreds of years of use through which history, memory, social action, and space intersected and contributed to its formation. Further, the traditional idea that Tibes was the seat of a powerful chiefdom (Curet 1992a, 1996; Curet and Oliver 1998; González Colón 1984; Rouse 1992; Siegel 1996) is perhaps not the only model we should be considering in our attempts to gain a better understanding of the site or the emergence of social “complexity” on the island. The chiefdom model implies that the site was the location in the region from which social change and power emanated and part of a linear evolutionary process. Rather, the site formed part of a broader relational system of socialities that must be considered before we can ultimately understand it within the broader contexts of organization, change, and development. To this end, our understanding of Tibes must take into account alternative forms of organization and social and cultural process that challenge traditional interpretive models regarding the development and organization of Tibes within the broader frameworks of Caribbean archaeology. Research in the present volume has begun to move toward this endeavor, not only by obtaining detailed data at multiple scales but also by being open to new lines of thinking that begin to question traditional models. Through this approach, we are trying to explain patterns by using multiple models in a recursive process wherein data are informed by theory at the same time that theory is reshaped by data. It is only through constantly questioning our assumptions, examining multiple lines of evidence, and rethinking our perspectives that we can form new interpretations, histories, and a better understanding of the past. This volume does not represent the final chapter of Tibes, but the beginning of a new frame of reference from which to push us forward in understanding its past and that of Puerto Rico.

References Cited

Abbott, R. Tucker 1974 American Seashells: The Marine Mollusca of the Atlantic and Pacific Coasts of North America. 2nd ed. Van Nostrand Reinhold, New York. Abruña, Fernando, Miguel A. Lugo López, and Raúl Pérez Escolar 1977 Los suelos de Puerto Rico. In Geovisión de Puerto Rico, edited by M. T. B. de Galiñanes, pp. 123–157. Editorial Universitaria, Universidad de Puerto Rico, Río Piedras. Adler, Michael A. 2002 The Ancestral Pueblo Community as Structure and Strategy. In Seeking the Center Place, edited by M. D. Varien and R. H. Wilshusen, pp. 25–39. University of Utah Press, Salt Lake City. Adler, Michael A., and Richard H. Wilshusen 1990 Large-Scale Integrative Facilities in Tribal Societies: Cross-Cultural and Southwestern US Examples. World Archaeology 22(2):133–146. Aitken, R. 1918 A Porto Rico Burial Cave. American Anthropologist 20:269–309. Alegría, Ricardo E. 1983 Ball Courts and Ceremonial Plazas in the West Indies. Yale University Publications in Anthropology No. 79. Department of Anthropology, Yale University, New Haven, Connecticut. Alvarado Zayas, Pedro 1981 La cerámica del centro ceremonial de Tibes: Estudio descriptivo. Unpublished Master’s thesis, Centro de Estudios Avanzados de Puerto Rico y el Caribe, San Juan, Puerto Rico. Ambrose, Stanley H. 1987 Chemical and Isotopic Techniques of Diet Reconstruction in Eastern North America. In Emergent Horticultural Economies of the Eastern Woodlands, ed-

288 / References Cited ited by William F. Keegan, pp. 87–107. Occasional Paper No. 7. Center for Archaeological Investigations, Southern Illinois University, Carbondale. 1990 Preparation and Characterization of Bone and Tooth Collagen for Stable Carbon and Nitrogen Isotope Analysis. Journal of Archaeological Science 17:430–451. 1993 Isotopic Analysis of Paleodiets: Methodological and Interpretive Considerations. In Investigation of Ancient Human Tissue: Chemical Analysis in Anthropology, edited by M. K. Sanford, pp. 59–130. Gordon and Breach, Amsterdam. 2000 Controlled Diet and Climate Experiments on Nitrogen Isotope Ratios of Rats. In Biogeochemical Approaches to Paleodietary Analysis, edited by Stanley H. Ambrose and M. Anne Katzenberg, pp. 243–259. Kluwer Academic/ Plenum, New York. Ambrose, Stanley H., Jane Buikstra, and Harold W. Krueger 2003 Status and Gender Differences in Diet in Mound 72, Cahokia, Revealed by Isotopic Analysis of Bone. Journal of Anthropological Archaeology 22:217–226. Ambrose, Stanley H., Brian M. Butler, Douglas B. Hanson, R. L. Hunter-Anderson, and Harold W. Krueger 1997 Stable Isotopic Analysis of Human Diet in the Marianas Archipelago, Western Pacific. American Journal of Physical Anthropology 104:343–361. Ambrose, Stanley H., and Lynette Norr 1993 Experimental Evidence for the Relationship of the Carbon Isotope Ratios of Whole Diet and Dietary Protein to Those of Bone Collagen and Carbonate. In Prehistoric Bone: Archaeology at the Molecular Level, edited by Joseph B. Lambert and Gisela Grupe, pp. 1–37. Springer-Verlag, New York. Ames, Kenneth M. 2004 Supposing Hunter-Gatherer Variability (book review essay). American Antiquity 69(2):364–374. Anderson, David G. 1996a Fluctuations between Simple and Complex Chiefdoms: Cycling in the Late Prehistoric Southeast. In Political Structure and Change in the Prehistoric Southeastern United States, edited by John F. Scarry, pp. 231–252. University Press of Florida, Gainesville. 1996b Chiefly Cycling and Large-Scale Abandonments as Viewed from the Savannah River Basin. In Political Structure and Change in the Prehistoric Southeastern United States, edited by John F. Scarry, pp. 150–191. University Press of Florida, Gainesville. Anderson Córdova, Karen F. 1990 Hispaniola and Puerto Rico: Indian Acculturation and Heterogeneity, 1492– 1550. Ph.D. dissertation, Yale University, New Haven, Connecticut. University Microfilms International, Ann Arbor. Appadurai, Arjun 1981 Gastro-Politics in Hindu South Asia. American Ethnologist 8(3):494–511. Ashley, Bob, Joanne Hollows, Steve Jones, and Ben Taylor 2004 Food and Cultural Studies. Routledge, London. Atran, Scott 1993 Itza Maya Tropical Agro-Forestry. Current Anthropology 34(5):633–689.

References Cited / 289 Aufderheide, A. C., and Conrado Rodríguez-Martín 1998 The Cambridge Encyclopedia of Human Paleopathology. Cambridge University Press, New York. Austin, Daniel F. 2004 Florida Ethnobotany. CRC Press, Boca Raton, Florida. Ayensu, Edward S. 1981 Medicinal Plants of the West Indies. Reference Publications, Algonac, Michigan. Ayes Suárez, Carlos 1989 Angostura: Un campamento arcaico temprano del Valle de Manataubón. Revista Universidad de América 1:24–37. Baetsen, S. 1999 Human Remains. In Archaeological Investigations on St. Martin (Lesser Antilles), edited by C. L. Hofman and M. L. P. Hoogland, pp. 249–250. Faculty of Archaeology, Leiden University, Leiden, the Netherlands. Barberena, R., and L. A. Borrero 2005 Stable Isotopes and Faunal Bones. Comments on Milner et al. (2004). Antiquity 79(303):191–195. Bass, William M. 1987 Human Osteology: A Laboratory and Field Manual. Missouri Archaeological Society. Basso, Ellen B. 1977 Introduction: The Status of Carib Ethnography. In Carib-Speaking Indians: Culture, Society, and Language, edited by Ellen B. Basso, pp. 9–22. Anthropological Papers of the University of Arizona No. 28. University of Arizona Press, Tucson. Baumforth, Douglas B., and Nyree Finlay 2008 Introduction: Archaeological Approaches to Lithic Production Skill and Craft Learning. Journal of Archaeological Method and Theory 15:1–27. Bean, L. J., and J. W. Lawton 1993 Some Explanations for the Rise of Cultural Complexity in Native California with Comments on Proto-Agriculture and Agriculture. In Before the Wilderness: Environmental Management by Native Californians, edited by T. C. Blackburn and K. Anderson, pp. 27–54. Ballena Press, Menlo, California. Bender, Margaret M., David A. Baerreis, and Raymond L. Steventon 1981 Further Light on Carbon Isotopes and Hopewell Agriculture. American Antiquity 46(2):346–353. Benedetti, María 1998 Earth and Spirit: Medicinal Plants and Healing Lore from Puerto Rico. Verde Luz, Cayey, Puerto Rico. Bequette, Kathryn 1991 Shipwrecks of St. Eustatius: A Preliminary Study. In Proceedings of the Thirteenth International Congress for Caribbean Archaeology, edited by E. N. Ayubi and J. B. Haviser, 787–800. Reports of the Archaeological-Anthropological Institute of the Netherlands Antilles No. 9. Curaçao. Bhaskar, S. N. 1973 Synopsis of Oral Pathology. CV Mosby, Saint Louis.

290 / References Cited Blanton, Richard E., Gary M. Feinman, Steve A. Kowalewski, and P. M. Peregrine 1996 A Dual-Processual Theory for the Evolution of Mesoamerican Civilization. Current Anthropology 37:1–14. Bocherens, Hervé, Dorotheé G. Drucker, Daniel Billiou, Marylène Patou-Mathis, and Bernard Vandermeersch 2005 Isotopic Evidence for Diet and Subsistence Pattern of the Saint-Césaire I Neanderthal: Review and Use of a Multi-Source Mixing Model. Journal of Human Evolution 49:71–87. Bocherens, Hervé, Marilyn L. Fogel, Noreen Tuross, and Melinda Zeder 1995 Trophic Structure and Climatic Information from Isotopic Signatures in Pleistocene Cave Fauna of Southern England. Journal of Archaeological Science 22:327–340. Boomert, Arie 2001 Saladoid Sociopolitical Organization. In Proceedings of the Eighteenth International Congress for Caribbean Archaeology, Vol. 2, edited by G. Richard, pp. 55–77. International Association for Caribbean Archaeology, Guadeloupe. Borghi, Alessandro, Lorenzo Appolonia, Laura Fiora, and Zoja Alessandro 2006 The Grey Marble of Porta Praetoria (Aosta, Italy), a Minero-Petrographic Characterisation and Provenance Determination. Periodico di Mineralogia 75:59–74. Bösl, Cornelia, Gisela Grupe, and Joris Peters 2006 A Late Neolithic Vertebrate Food Web Based on Stable Isotope Analyses. International Journal of Osteoarchaeology 16(4):296–315. Bourdieu, Pierre 1977 Outline of a Theory of Practice. Translated by R. Nice. Cambridge Studies in Social and Cultural Anthropology. Cambridge University Press, Cambridge. Breiner, Sheldon 1973 Applications Manual for Portable Magnetometers. GeoMetrics, Sunnyvale, California. Brenna, J. Thomas, Thomas N. Corso, Herbert J. Tobias, and Richard J. Caimi 1997 High-Precision Continuous-Flow Isotope Ratio Mass Spectrometry. Mass Spectrometry Reviews 16(5):227–258. Bright, Alistair 2003 Spatial Dynamics and Social Development in the Northern Lesser Antilles: A Pilot Study Based on Settlement Structure at the Site of Anse à la Gourde, Guadeloupe. Unpublished dissertation, Department of Archaeology, Leiden University, Leiden, the Netherlands. Brink, Jack W., and Bob Dawe 2003 Hot Rocks as Scarce Resources: The Use, Re-Use and Abandonment of Heating Stones at Head-Smashed-In Buffalo Jump. Plains Anthropologist 48:85–104. Brumfiel, Elizabeth M. 1992 Distinguished Lecture in Archaeology: Breaking and Entering the Ecosystem—Gender, Class, and Action Steal the Show. American Anthropologist 94:551–567.

References Cited / 291 Brush, Stephen B. 2004 Farmers’ Bounty: Locating Crop Diversity in the Contemporary World. Yale University Press, New Haven, Connecticut. Budinoff, Linda C. 1991 An Osteological Analysis of Human Burials Recovered from Maisabel: An Early Ceramic Site on the North Coast of Puerto Rico. In Proceedings of the Twelfth International Congress for Caribbean Archaeology, edited by L. S. Robinson, pp. 117–133. Association Internationale d’Archeologie de la Caraibe, Martinique. Buikstra, Jane E. 1992 Diet and Disease in Late Prehistory. In Disease and Demography in the Americas, edited by John W. Verano and Douglas H. Ubelaker, pp. 87–101. Smithsonian Institution Press, Washington, D.C. 2006 Bioarchaeology: The Contextual Analysis of Human Remains. Elsevier Academic, San Diego, California. Buikstra, Jane E., and Douglas H. Ubelaker (editors) 1994 Standards for Data Collection from Human Skeletal Remains: Proceedings of a Seminar at the Field Museum of Natural History. Arkansas Archeological Survey Research Series No. 44. Arkansas Archeological Survey, Fayetteville. Carbone, Victor 1980 Cultural Resource Assessment of the Río Grande de Loíza Project Area and Environs. Interagency Archaeological Services, U.S. Department of the Interior, National Park Service, Southeast Regional Office, Atlanta. Carroll, Lynda 1999 Communities and Other Social Actors: Rethinking Commodities and Consumption in Global Historical Archaeology. International Journal of Historical Archaeology 3(3):131–136. Chang, K. C. (editor) 1968 Settlement Archaeology. National Press Books, Palo Alto, California. Chanlatte Baik, Luis A. 1976 Investigaciones arqueológicas en Guayanilla, Puerto Rico. Museo del Hombre Dominicano, Santo Domingo. 1981 La Hueca y Sorcé (Vieques, Puerto Rico): Primeras migraciones agroalfareras antillanas. Published by the author, Santo Domingo, Dominican Republic. 1995 La presencia huecoide en Hacienda Grande, Loíza. In Proceedings of the Fifteenth International Congress for Caribbean Archaeology, edited by R. E. Alegría and M. Rodríguez, pp. 501–510. Centro de Estudios Avanzados de Puerto Rico y el Caribe, San Juan. Chanlatte Baik, Luis A., and Yvonne M. Narganes Storde 1980 La Hueca, Vieques: Nuevo complejo cultural agroalfarero en la arqueología antillana. Proceedings of the International Congress for the Study of Precolumbian Culture of Lesser Antilles 8:501–523. Arizona State University, Tempe. 1983 Vieques-Puerto Rico: Asiento de una nueva cultura aborigen antillana. Published by the authors, Santo Domingo, Dominican Republic. 1986 Proceso y desarrollo de los primeros pobladores de Puerto Rico y las Antillas. Published by the authors, Santo Domingo, Dominican Republic.

292 / References Cited Chisholm, Brian S., D. Erle Nelson, K. A. Hobson, Henry P. Schwarcz, and M. Knyf 1983 Carbon Isotope Measurement Techniques for Bone Collagen: Notes for the Archaeologist. Journal of Archaeological Science London 10(4):355–360. Chisholm, Brian S., D. Erle Nelson, and Henry P. Schwarcz 1982 Stable-Carbon Isotope Ratios as a Measure of Marine versus Terrestrial Protein in Ancient Diets. Science 216(4550):1131–1132. Clark, Anthony 1990 Seeing Beneath the Soil: Prospecting Methods in Archaeology. B. T. Batsford, London. Clark, Jeffrey J., Jeff B. Walker, and Reniel Rodríguez Ramos 2003 Depositional History and Evolution of the Paso del Indio Site, Vega Baja, Puerto Rico. Geoarchaeology: An International Journal 18:625–648. Cohen, Anthony P. 1985 The Symbolic Construction of Community. Routledge, New York. Cohen, Mark N., and George J. Armelagos 1984 Paleopathology at the Origins of Agriculture. Academic Press, Orlando, Florida. Cotterell, Brian, and Johan Kamminga 1990 Mechanics of Pre-Industrial Technology. Cambridge University Press, Cambridge. Cowgill, George L. 1993 Distinguished Lecture in Archaeology: Beyond Criticizing New Archaeology. American Anthropologist 95:551–573. Craig, Harmon 1957 Isotopic Standards for Carbon and Oxygen and Correction Factors for MassSpectrometric Analysis of Carbon Dioxide. Geochimica et Cosmochimica Acta 12:133–149. Crespo-Torres, Edwin F. 1987 Estudio antropofísico de las entierros humanos procedentes del parque de baseball del pueblo de Comerio, Puerto Rico. Manuscript on file, Instituto de Cultura Puertorriqueña, San Juan. 1991 Informe preliminar sobre los enterramientos humanos en el yacimiento de Punta Candelero, Puerto Rico. In Proceedings of the Thirteenth International Congress for Caribbean Archaeology (Part 2), edited by E. N. Ayubi and J. B. Haviser, pp. 840–853. Reports of the Archaeological-Anthropological Institute of the Netherlands Antilles No. 9. Curaçao. 1994 Dental Analysis of Human Burials Recovered from Punta Candelero: A Prehistoric Site on the Southeast Coast of Puerto Rico. Unpublished Master’s thesis, Department of Anthropology, Arizona State University, Tempe. 1998 Osteobiografía de los restos humanos procedentes del centro ceremonial indígena de Tibes, Ponce, Puerto Rico. Submitted to the Municipio Autónomo de Ponce, Puerto Rico. 2000a Estudio comparativo biocultural entre dos poblaciones prehistóricas en la isla de Puerto Rico: Punta Candelero y Paso del Indio. Unpublished Ph.D. dissertation, Instituto de Investigaciones Antropológicas, Universidad Nacional Autónoma de México, México City. 2000b Reporte preliminar del análisis osteológico de los restos humanos procedentes del sitio arqueológico de Lujan, Vieques, Puerto Rico. Manuscript submitted to Virginia Rivera (Principal Investigator), CRM Project.

References Cited / 293 2001a Informe final sobre los entierros descubiertos durante las excavaciones arqueológicas que se llevaron acabo en la comunidad Esperanza en el municipio de Vieques, Puerto Rico, como parte del Proyecto de Mejoras Sanitarias, de dicha comunidad. Manuscript submitted to Marlene Ramos (Principal Investigator), CRM Project. 2001b Proyecto arqueológico Paso del Indio, fase III, estudio osteológico de los restos humanos. Manuscript on file, Paso del Indio Archaeological Project, Puerto Rico Highway and Transportation Authority, Commonwealth of Puerto Rico, San Juan. 2002 Nuevas interpretaciones en torno a las creencias sobre la muerte y prácticas funerarias de los indios de Boriquén. Revista del Instituto de Cultura Puertorriqueña 3(5):83–94. 2004 Análisis osteológico de restos humanos procedentes de Maruca: Un sitio precerámico al sur de la isla de Puerto Rico (Part VII). In Excavaciones en el yacimiento arcaico de Maruca, Ponce, Puerto Rico, by M. Rodríguez. Final report submitted to Consejo para la Protección del Patrimonio Arqueológico Terrestre de Puerto Rico and Programa de Arqueología y Etnohistoria del Instituto de Cultura Puertorriqueña, San Juan. 2005 La cultura huecoide y su conexión con la introducción de la práctica de la deformación cefálica intencional en las Antillas. In Cultura La Hueca, pp. 57–65. Museo de Historia, Antropología y Arte, Universidad de Puerto Rico, Recinto de Río Piedras. Crock, John G. 2000 Interisland Interaction and the Development of Chiefdoms in the Eastern Caribbean. Unpublished Ph.D. dissertation, Department of Anthropology, University of Pittsburgh, Pittsburgh. Crumley, Carol L. 1995 Heterarchy and the Analysis of Complex Societies. In Heterarchy and the Analysis of Complex Societies, edited by R. M. Ehrenreich, C. L. Crumley, and J. E. Levy, pp. 1–6. Archeological Papers of the American Anthropological Association No. 6. American Anthropological Association, Washington, D.C. Curet, L. Antonio 1992a The Development of Chiefdoms in the Greater Antilles: A Regional Study of the Valley of Maunabo, Puerto Rico. Unpublished Ph.D. dissertation, Department of Anthropology, Arizona State University, Tempe. 1992b House Structure and Cultural Change in the Caribbean: Three Case Studies from Puerto Rico. Latin American Antiquity 3:160–174. 1993 Prehistoric Demographic Changes in the Valley of Maunabo: A Preliminary Report. In Proceedings of the Fourteenth International Congress for Caribbean Archaeology, edited by A. Cummins and P. King, pp. 11–24. International Association for Caribbean Archaeology, Barbados. 1996 Ideology, Chiefly Power and Material Culture: An Example from the Greater Antilles. Latin American Antiquity 7:114–131. 2002 The Chief Is Dead, Long Live . . . Who? Descent and Succession in the Protohistoric Chiefdoms of the Greater Antilles. Ethnohistory 49(2):259–280. 2003 Issues on the Diversity and Emergence of Middle-Range Societies of the Ancient Caribbean: A Critique. Journal of Archaeological Research 11:1–42.

294 / References Cited 2004

Island Archaeology and Units of Analysis in the Study of Ancient Caribbean Societies. In Voyages of Discovery: The Archaeology of Islands, edited by S. M. Fitzpatrick, pp. 187–201. Praeger, Westport, Connecticut. 2005 Caribbean Paleodemography: Population, Culture History, and Sociopolitical Processes in Ancient Puerto Rico. The University of Alabama Press, Tuscaloosa. 2006 Description of Excavation Units of the Proyecto Arqueológico del Centro Ceremonial de Tibes (1995–2003). Manuscript on file, Field Museum of Natural History, Chicago. Curet, L. Antonio, and Lee A. Newsom 1998 Report on the 1996–1997 Research Conducted by the Archaeological Project of the Civic-Ceremonial Center of Tibes, Ponce, Puerto Rico. Latin American Archaeology Program, Heinz Family Foundation. On file, Field Museum of Natural History, Chicago. Curet, L. Antonio, Lee A. Newsom, and Susan D. deFrance 2006 Cultural Continuity and Discontinuity in the Social History of Ancient Puerto Rico: The Case of the Ceremonial Center of Tibes. Journal of Field Archaeology 31(1):23–39. Curet, L. Antonio, Lee A. Newsom, and Daniel Welch 2003 Space and Time in the Civic-Ceremonial Center of Tibes, Ponce, Puerto Rico. In Proceedings of the Nineteenth International Congress for Caribbean Archaeology, edited by L. Alofs and R. A. C. F. Dijkhoff, pp. 142–155. Publication of the Museo Arqueológico Aruba, Vol. 9(2), Aruba. Curet, L. Antonio, and José R. Oliver 1998 Mortuary Practices, Social Developments, and Ideology in Precolumbian Puerto Rico. Latin American Antiquity 9:217–239. Curet, L. Antonio, Joshua M. Torres, and Miguel Rodríguez 2004 Political and Social History of Eastern Puerto Rico: The Ceramic Age. In Late Ceramic Age Societies in the Eastern Caribbean, edited by A. Delpuech and C. Hofman, pp. 59–85. BAR International Series S1273, Paris Monographs in American Archaeology 14. Archaeopress, Oxford. Curet, L. Antonio, Daniel Welch, W. Derek Hamilton, and Lee A. Newsom 2005 In Search of the Lost Features: A Progress Report of a Geophysical Study at the Civic-Ceremonial Center of Tibes, Puerto Rico. In Proceedings of the Twentieth International Congress for Caribbean Archaeology, edited by Glenis Tavares María and Manuel García Arévalo, pp. 65–74. Museo del Hombre Dominicano and Fundación García Arévalo, Santo Domingo. Curtis, J. H., M. Brenner, and D. A. Hodell 2001 Climate Change in the Circum-Caribbean (Late Pleistocene to Present) and Implications for Regional Biogeography. In Biogeography of the West Indies: Patterns and Perspectives, edited by C. A. Woods and F. E. Sergile, pp. 35–54. CRC Press, Boca Raton, Florida. Danforth, Marie E. 1999 Nutrition and Politics in Prehistory. Annual Review of Anthropology 28(1):1–25. David, Andrew, Neil Linford, and Paul Linford 2008 Geophysical Survey in Archaeological Field Evaluation. English Heritage, Swindon, U.K.

References Cited / 295 Davies, C. E., and D. Moss 2002 EUNIS Habitat Classification. Final Report to the European Topic Centre on Nature Protection and Biodiversity, European Environment Agency, Paris. Deagan, Kathleen 1983 Spanish Florida as Part of the Caribbean Colonial Sphere. In Proceedings of the Ninth International Congress for the Study of the Pre-Columbian Cultures of the Lesser Antilles, edited by L. Allaire, pp. 419–429. Centre de Recherche Caraïbes, Université de Montréal. 2004 Reconsidering Taíno Social Dynamics after Spanish Conquest: Gender and Class in Culture Contact Studies. American Antiquity 69(4):597–626. deFrance, Susan D. 1988 Zooarchaeological Investigations of Subsistence Strategies at the Maisabel Site, Puerto Rico. Unpublished Master’s paper, Department of Anthropology, University of Florida, Gainesville. 1989 Saladoid and Ostionoid Subsistence Adaptations: Zooarchaeological Data from a Coastal Occupation on Puerto Rico. In Early Ceramic Population Lifeways and Adaptive Strategies in the Caribbean, edited by P. E. Siegel, pp. 57–77. BAR International Series 506. British Archaeological Reports, Oxford. 1990 Zooarchaeological Investigations of an Early Ceramic Age Frontier Community in the Caribbean, Maisabel Site, Puerto Rico. Antropológica 73–74:3–180. 2009 Chiefly Fare or Who’s Feeding the Cacique? Equality in Animal Use at the Tibes Ceremonial Center, Puerto Rico. In Anthropological Approaches to Zooarchaeology: Colonialism, Complexity, and Animal Transformations, edited by D. Campana, A. Choyke, P. Crabtree, S. deFrance, and J. Lev-Tov. Oxbow Books, Oxford. deFrance, Susan D., William F. Keegan, and Lee A. Newsom 1996 The Archaeobotanical, Bone Isotope, and Zooarchaeological Records from Caribbean Sites in Comparative Perspective. In Case Studies in Environmental Archaeology, edited by E. J. Reitz, L. A. Newsom, and S. J. Scudder, pp. 289– 304. Plenum Press, New York. deFrance, Susan D., and Michelle J. LeFebvre 2009 New Records of Guinea Pig Remains from the Lesser Antilles. Poster presented at the 74th Annual Meeting of the Society for American Archaeology, Atlanta. deFrance, Susan D., and Lee A. Newsom 2005 The Status of Paleoethnobiological Research on Puerto Rico. In Ancient Borinquen: Archaeology and Ethnohistory of Native Puerto Rico, edited by Peter E. Siegel, pp. 122–184. The University of Alabama Press, Tuscaloosa. De Hostos, Adolfo 1941 Notes on West Indian Hydrography and Its Relation to Prehistoric Migrations. In Anthropological Papers: Papers Based Principally on Studies of the Prehistoric Archaeology and Ethnology of the Greater Antilles, edited by A. De Hostos, pp. 30–53. Office of the Historian, Government of Puerto Rico, San Juan. Delwiche, C. C., P. J. Zinke, C. M. Johnson, and R. A. Virgina 1979 Nitrogen Isotope Distribution as a Presumptive Indicator of Nitrogen Fixation. Botanical Gazette 140:65–69.

296 / References Cited DeNiro, Michael J. 1985 Postmortem Preservation and Alteration of In Vivo Bone Collagen Isotope Ratios in Relation to Palaeodietary Reconstruction. Nature 317:806–809. 1987 Stable Isotopy and Archaeology. American Scientist 75:182–191. 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. Divakarla, Pawan, and Emir J. Macari 1998 Stability Analysis of Slopes in Western Puerto Rico. In Stability of Natural Slopes in the Coastal Plain, edited by Timothy R. Martin, pp. 59–68. American Society of Civil Engineers Geotechnical Special Publication No. 77. American Society of Civil Engineers, Reston, Virginia. Dobkin de Rios, Marlene 1984 Hallucinogens: Cross-Cultural Perspectives. University of New Mexico Press, Albuquerque. Drew, R. 2005 A Survey of Pre-Columbian Human Remains from Puerto Rico. In Proceedings of the Twentieth International Congress for Caribbean Archaeology, edited by Glenis Tavares María and Manuel García Arévalo, pp. 523–534. Museo del Hombre Dominicano and Fundación García Arévalo, Santo Domingo. Drucker, Dorothée G., and Dominique Henry-Gambier 2005 Determination of the Dietary Habits of a Magdalenian Woman from SaintGermain-la-Riviére in Southwestern France Using Stable Isotopes. Journal of Human Evolution 49:19–35. Duke, James A. 1992 Handbook of Edible Weeds. CRC Press, Boca Raton, Florida. Dunn, Oliver, and James E. Kelley Jr. (translators) 1989 The Diario of Christopher Columbus’s First Voyage to America, 1492–1493, Abstracted by Fray Bartolomé de las Casas. University of Oklahoma Press, Norman. Durand, Jean-François, and Henry Petitjean Roget 1991 A propos d’un collier funeraire a Morel, Guadeloupe: Les huecoide sont-ils un mythe? In Proceedings of the Twelfth International Congress for Caribbean Archaeology, edited by L. S. Robinson, pp. 53–72. Association Internationale d’Archeologie de la Caraibe, Martinique. Earle, Timothy K. 1997 How Chiefs Come to Power: The Political Economy in Prehistory. Stanford University Press, Stanford, California. El-Najjar, M. Y. 1977 The Distribution of Human Skeletal Material in the Continental United States. American Journal of Physical Anthropology 46:507–512. Ericson, Jonathon E., and Michael D. Glascock 2004 Subsource Characterization: Obsidian Utilization of Subsources of the Coso Volcanic Field, Coso Junction, California, USA. Geoarchaeology 19:779–805. Espenshade, Christopher T. 2000 Reconstructing Household Vessel Assemblages and Site Duration at an Early

References Cited / 297 Ostionoid Site from South-Central Puerto Rico. Journal of Caribbean Archaeology 1(1):1–22. Espenshade, Christopher T., J. Foss, and Joseph W. Joseph 1987 Data Recovery Excavations at Site PO-21, Cerrillos River Valley, Puerto Rico. Garrow and Associates, Inc., Atlanta. Submitted to the U.S. Army Corps of Engineers, Jacksonville District, Jacksonville, Florida. Espenshade, Christopher T., Peter E. Siegel, J. Loubser, and H. B. Matternes 2007 Detailed Phase II Archaeological Testing and Data Recovery Planning, Site PO-29, Portugues Dam and Pool, Puerto Rico. New South Associates, Inc., Stone Mountain, Georgia. Submitted to the U.S. Army Corps of Engineers, Jacksonville District, Jacksonville, Florida. Espenshade, Christopher T., and Stacey Young (compilers) 2008 End of Field Progress Report: Phase III Investigations of Site PO-29, Municipio de Ponce, Puerto Rico. New South Associates, Inc., Stone Mountain, Georgia. Submitted to the U.S. Army Corps of Engineers, Jacksonville District, Jacksonville, Florida. Ewel, John, and Jacob L. Whitmore 1973 Las zonas de vida de Puerto Rico y las Islas Vírgenes americanas, una sinópsis. Instituto de Dasonomía Tropical, U.S. Department of Agriculture, Río Piedras, Puerto Rico. Fairbanks, Charles H., and Rochelle A. Marrinan 1983 The Puerto Real Project, Haiti. In Proceedings of the Ninth International Congress for the Study of the Pre-Columbian Cultures of the Lesser Antilles, edited by L. Allaire, pp. 409–418. Centre de Recherche Caraïbes, Université de Montréal. Farb, Peter, and George J. Armelagos 1983 Consuming Passions: The Anthropology of Eating. Washington Square Press, New York. Feinman, Gary M. 1995 The Emergence of Inequality: A Focus on Strategies and Processes. In Foundations of Social Inequality, edited by T. D. Price and G. M. Feinman, pp. 255– 279. Plenum Press, New York. Feinman, Gary M., Ken G. Lightfoot, and Steve Upham 2000 Political Hierarchies and Organizational Strategies in the Puebloan Southwest. American Antiquity 65:449–470. Fenner, Michael, and Ken Thompson 2005 The Ecology of Seeds. 2nd ed. Cambridge University Press, Cambridge. Fernández Méndez, Eugenio 1972 Arte y mitología de los indios Taínos de las Antillas Mayores. Ediciones El Cemí, San Juan, Puerto Rico. Fernández de Oviedo y Valdés, Gonzalo 1851– Historia general y natural de las Indias, islas y tierra firme del mar océano, 1855 4 vols. Edited by J. Amador de los Ríos. Real Academia de la Historia, [1535] Madrid. 1959a Natural History of the West Indies. Translated and edited by S. A. Stoudemire. [1526] University of North Carolina Studies in the Romance Languages and Literature No. 32. University of North Carolina Press, Chapel Hill.

298 / References Cited 1959b Historia general y natural de las Indias. In Biblioteca de autores españoles, vols. 117–122. Ediciones Atlas, Madrid. 1975 The Conquest and Settlement of the Island of Boriquen or Puerto Rico. Translated and edited by Daymond Turner. Cardavon Press, Avon, Connecticut. Fewkes, Jesse W. 1970 The Aborigines of Porto Rico and Neighboring Islands. Johnson Reprint Corp., New York. Originally published 1907, Annual Report of the Bureau of American Ethnology for 1903–1904, No. 25. Smithsonian Institution, Washington, D.C. Frayer, David 1988 Auditory Exostoses and Evidence for Fishing at Vlasac. Current Anthropology 29:346–349. Fry, Brian 1991 Stable Isotope Diagrams of Freshwater Food Webs. Ecology 72(6):2293–2297. 2006 Stable Isotope Ecology. Springer, New York. Fry, Brian, Ruth Lutes, Mark Northam, and Patrick L. Parker 1982 A 13C/12C Comparison of Food Webs in Caribbean Seagrass Meadows and Coral Reefs. Aquatic Botany 14(4):389–398. Gaffney, Chris, and John Gater 2003 Revealing the Buried Past: Geophysics for Archaeologists. Tempus, Stroud, Gloucestershire, Great Britain. García Arévalo, Miguel 1990 El juego de pelota taíno y su importancia comercial. In Proceedings of the Eleventh International Congress for Caribbean Archaeology, edited by A. G. Pantel Tekakis, I. Vargas Arenas, and M. Sanoja Obediente, pp. 178–186. La Fundación Arqueológica, Antropológica e Histórica de Puerto Rico, San Juan. García Goyco, Osvaldo 1983 Influencias mayas y aztecas en los Taínos de las Antillas Mayores. Ediciones Xibalbay, San Juan, Puerto Rico. Garrow, Patrick H., Charles H. McNutt Jr., Guy G. Weaver, and José Oliver 1995 La Iglesia de Maraguez (PO-39): Investigation of a Local Ceremonial Center on the Cerrillos River Valley, Ponce, Puerto Rico. Garrow and Associates, Inc., Atlanta. Geertz, Clifford 1980 Negara: The Theatre State in Nineteenth-Century Bali. Princeton University Press, Princeton, New Jersey. Gero, Joan M. 1991 Genderlithics: Women’s Roles in Stone Tool Production. In Engendering Archaeology: Women in Prehistory, edited by Joan M. Gero and Margaret Conkey, pp. 163–193. Basil Blackwell, Oxford. Giddens, Anthony 1984 The Constitution of Society. University of California Press, Berkeley. Gierbolini, Roberto E. 1979 Soil Survey of Ponce Area of Southern Puerto Rico. U.S. Department of Agriculture, Soil Conservation Service, Washington, D.C., in cooperation with the University of Puerto Rico, College of Agricultural Sciences.

References Cited / 299 Gilmore, R. Grant, III 2008 Geophysics and Volcanic Islands: Resistivity and Gradiometry on St. Eustatius. In Archaeology and Geoinformatics: Case Studies from the Caribbean, edited by Basil A. Reid, pp. 170–183. The University of Alabama Press, Tuscaloosa. González, Juan, and Luis A. Rodríguez 1985 Sitio Caracoles: Ponce, Puerto Rico. Proyecto de investigación. Submitted to Programa de Arqueología, Instituto de Cultura Puertorriqueña, San Juan. González Colón, Juan 1984 Tibes: Un centro ceremonial indígena. Unpublished Master’s thesis, Centro de Estudios Avanzados de Puerto Rico y el Caribe, San Juan, Puerto Rico. Goodman, Alan H., R. Brooke Thomas, Alan C. Swedlund, and George J. Armelagos 1988 Biocultural Perspectives on Stress in Prehistoric, Historical, and Contemporary Population Research. American Journal of Physical Anthropology 31(S9):169–202. Goodwin, R. Christopher, and Jeff B. Walker 1975 The Excavation of an Early Taino Site in Puerto Rico. Inter-American University Press, San Juan, Puerto Rico. Gorlin, Robert J., and Henry M. Goldman (editors) 1970 Thoma’s Patología Oral. Editorial Salvat, México City. Gray, John 1999 Open Spaces and Dwelling Places: Being at Home on Hill Farms in the Scottish Borders. In The Anthropology of Space and Place, edited by S. M. Low and D. Lawrence-Zúñiga, pp. 224–244. Blackwell, Malden, Massachusetts. Greenough, John D., M. P. Gorton, and Leanne M. Mallory-Greenough 2001 The Major- and Trace-Element Whole-Rock Fingerprints of Egyptian Basalts and the Provenance of Egyptian Artifacts. Geoarchaeology 16:763–784. Guarch Delmonte, José 1972 Excavaciones en el extremo oriental de Cuba. Serie Arqueológica No. 1. Academia de Ciencias de Cuba, Instituto de Arqueológica, Havana. Guita, J. L. 1984 Oral Pathology. Williams and Wilkins, Baltimore. Hadden, Carla J. 2007 Invertebrate Resource Use at the Tibes Ceremonial Center. Undergraduate Honors thesis, Department of Anthropology, University of Florida, Gainesville. On file at the University of Florida, Department of Anthropology, and on library microfiche at the University of Florida. Haglund, William D., and Marcela H. Sorg (editors) 1997 Forensic Taphonomy: The Postmortem Fate of Human Remains. CRC Press, Boca Raton, Florida. 2002 Advances in Forensic Taphonomy: Method, Theory, and Archaeological Perspectives. CRC Press, Boca Raton, Florida. Hale, H. Stephen, Irvy R. Quitmyer, and Sylvia Scudder 1987 Methods for Estimating Edible Meat Weights for Faunal Remains from Sites in the Southeastern United States. Manuscript on file, Environmental Archaeology Laboratory, Florida Museum of Natural History, University of Florida, Gainesville.

300 / References Cited Hale, H. Stephen, and Karen Jo Walker 1986 Allometric Regression of Osteichthyes Sample. Manuscript on file, Environmental Archaeology Laboratory, Florida Museum of Natural History, University of Florida, Gainesville. Hall, Robert L. 1967 Those Late Corn Dates: Isotopic Fractionation as a Source of Error in Carbon-14 Dates. Michigan Archaeologist 13(4):171–180. Harris, Marvin 1985 Good to Eat: The Sacred Cow and the Abominable Pig. Simon and Schuster, New York. Hastorf, Christine A., and Sissel Johannessen 1993 Pre-Hispanic Political Change and the Role of Maize in the Central Andes of Peru. American Anthropologist 95(1):115–138. Hawkins, Alicia L., and Maxine R. Kleindienst 2002 Lithic Raw Material Usage during the Middle Stone Age at Dalkleh Oasis, Egypt. Geoarchaeology 17:601–624. Hayden, Brian 1995 Pathways to Power: Principles for Creating Socioeconomic Inequalities. In Foundations of Social Inequality, edited by T. D. Price and G. M. Feinman, pp. 15–86. Plenum Press, New York. Heckenberger, Michael J. 2005 The Ecology of Power. Routledge, New York. Heckenberger, Michael J., Afukaka Kuikuro, Urissapáq Tabata Kuikuro, J. Christian Russell, Morgan Schmidt, Carlos Fausto, and Bruna Franchetto 2003 Amazonia 1492: Pristine Forest or Cultural Parkland? Science 301:1710–1714. Hedges, Robert E. M., and Linda M. Reynard 2007 Nitrogen Isotopes and the Trophic Level of Humans in Archaeology. Journal of Archaeological Science 34(8):1240–1251. Hegmon, Michelle 1989 Social Integration and Architecture. In The Architecture of Social Integration in Prehistoric Pueblos, edited by W. D. Lipe and M. Hegmon, pp. 5–15. Crow Canyon Archaeological Center, Cortez, Colorado. 2002 Concepts of Community in Archaeological Research. In Seeking the Center Place, edited by M. D. Varien and R. H. Wilshusen, pp. 263–279. University of Utah Press, Salt Lake City. Heidegger, Martin 1977 Basic Writings. Harper and Row, New York. Heimmer, Don H., and Steven L. De Vore 1995 Near-Surface, High Resolution Geophysical Methods for Cultural Resource Management and Archeological Investigations. U.S. Department of the Interior, National Park Service, Rocky Mountain Regional Office, Denver. Hirsch, Eric 1995 Introduction—Landscape: Between Space and Place. In The Anthropology of Landscape: Perspectives on Place and Space, edited by E. Hirsch and M. O’Hanlon, pp. 1–30. Clarendon Press, Oxford. Hirth, Kenneth G. 1996 Political Economy and Archaeology: Perspectives on Exchange and Production. Journal of Archaeological Research 4(3):203–239.

References Cited / 301 Hofman, Corinne L. 1995 Inferring Inter-Island Relationships from Ceramic Style: A View from the Leeward Islands. In Proceedings of the Fifteenth International Congress for Caribbean Archaeology, edited by R. E. Alegría and M. Rodríguez, pp. 233–241. Centro de Estudios Avanzados de Puerto Rico y el Caribe, San Juan, Puerto Rico. Hofman, Corinne L., and Menno L. P. Hoogland (editors) 1999 Archaeological Investigations on St. Martin (Lesser Antilles). Faculty of Archaeology, Leiden University, Leiden, the Netherlands. Hofman, Corinne L., Menno L. P. Hoogland, and Annelou L. van Gijn (editors) 2008 Crossing the Borders: New Methods and Techniques in the Study of Archaeological Materials from the Caribbean. The University of Alabama Press, Tuscaloosa. Honychurch, Penelope N. 1986 Caribbean Wild Plants and Their Uses. Macmillan, London. Hoogland, Menno L. P. 1996 In Search of the Native Population of Pre-Columbian Saba (400–1450 A.D.), Part Two: Settlement Sites in Their Natural and Social Environment. Unpublished Ph.D. dissertation, Department of Archaeology, Leiden University, Leiden, the Netherlands. Hoshower, Lisa M., Jane A. Buikstra, Paul S. Goldstein, and Ann D. Webster 1995 Artificial Cranial Deformation at the Omo M10 Site: A Tiawanaku Complex from the Moquegua Valley, Peru. Latin American Antiquity 6(2):145–164. Ingold, Tim 1993 The Temporality of the Landscape. World Archaeology 25:152–174. Inomata, Takeshi, and Lawrence S. Coben 2006 Scene I: Overture—Invitation to Archaeological Theater. In Archaeologies of Performance: Theaters of Power, Community and Politics, edited by T. Inomata and L. S. Coben, pp. 11–46. Altamira Press, Lanham, Maryland. Irish, Joseph D., and Christy G. Turner 1987 More Lingual Surface Attrition of the Maxillary Anterior Teeth in American Indians: Prehistoric Panamanians. American Journal of Physical Anthropology 73:209–213. Jane, Cecil 1988 The Voyage of Christopher Columbus. Argonaut Press, London. Johnson, Allen W., and Timothy Earle 1987 The Evolution of Human Societies: From Foraging Group to Agrarian State. Stanford University Press, Stanford, California. Keegan, William F. 1985 Dynamic Horticulturalists: Population Expansion in the Prehistoric Bahamas. Ph.D. dissertation, University of California, Los Angeles. University Microfilms, Ann Arbor. 1991 An Anthropological Evaluation of Taíno Kinship. In Proceedings of the Thirteenth International Congress for Caribbean Archaeology, edited by E. N. Ayubi and J. B. Haviser, pp. 437–445. Reports of the ArchaeologicalAnthropological Institute of the Netherlands Antilles No. 9. Curaçao. 1992 The People Who Discovered Columbus. University Press of Florida, Gainesville.

302 / References Cited 1996

West Indian Archaeology. 2: After Columbus. Journal of Archaeological Research 4(4):265–294. 1997a Bahamian Archaeology. Media Publishing, Nassau, Bahamas. 1997b “No Man [or Woman] Is an Island”: Elements of Taíno Social Organization. In The Indigenous People of the Caribbean, edited by S. M. Wilson, pp. 111– 117. University Press of Florida, Gainesville. 2000 West Indian Archaeology. 3: Ceramic Age. Journal of Archaeological Research 8(2)135–167. 2001 Archaeological Investigations at Ile à Rat, Haiti: Avoid the -oid. In Proceedings of the Eighteenth International Congress for Caribbean Archaeology, Vol. 2, edited by G. Richard, pp. 233–239. International Association for Caribbean Archaeology, Guadeloupe. 2004 Islands of Chaos. In Late Ceramic Age Societies in the Eastern Caribbean, edited by A. Delpuech and C. Hoffman, pp. 33–46. BAR International Series 1273. Archaeopress, Oxford. 2009 Central Plaza Burials in Saladoid Puerto Rico: An Alternative Perspective. Latin American Antiquity 20(2):1–11. Keegan, William F., and Michael J. DeNiro 1988 Stable Carbon- and Nitrogen-Isotope Ratios of Bone Collagen Used to Study Corral-Reef and Terrestrial Components of Prehistoric Bahamian Diet. American Antiquity 53(2):320–336. Keegan, William F., and Morgan D. Maclachlan 1989 The Evolution of Avunculocal Chiefdoms. American Anthropologist 91:613– 630. Keegan, William F., Morgan D. Maclachlan, and B. Byrne 1998 Social Foundations of Taino Caciques. In Chiefdoms and Chieftaincy in the Americas, edited by E. M. Redmond, pp. 217–244. University Press of Florida, Gainesville. Keegan, William F., Roger W. Portell, and John Slapcinsky 2003 Changes in Invertebrate Taxa at Two Pre-Columbian Sites in Southwestern Jamaica, AD 800–1500. Journal of Archaeological Science 3012:1607–1617. Keeling, Charles D., Wim G. Mook, and Pieter P. Tans 1979 Recent Trends in the 13C/12C Ratio of Atmospheric Carbon Dioxide. Nature 277:121–122. Kelly, Robert L. 1995 Egalitarian and Nonegalitarian Hunter-Gatherers. In The Foraging Spectrum: Diversity in Hunter-Gatherer Lifeways, pp. 293–331. Smithsonian Institution Press, Washington, D.C. Kendall, M. S., C. R. Cruer, K. R. Buja, J. D. Christensen, M. Finkbeiner, R. A. Warner, and M. E. Monaco 2001 Methods Used to Map the Benthic Habitats of Puerto Rico and the U.S. Virgin Islands. NOAA Technical Memorandum NOS NCCOS CCMA 152. Kennedy, Kenneth A. R. 1989 Skeletal Markers of Occupational Stress. In Reconstruction of Life from the Skeleton, edited by M. Y. Iscan and K. A. R. Kennedy, pp. 129–160. Alan R. Liss, New York. Kesseler, Rob, and Wolfgang Stuppy 2006 Seeds: Time Capsules of Life. Firefly Books, Buffalo, New York.

References Cited / 303 Kirch, Patrick V., and Sharyn Jones O’Day 2003 New Archaeological Insights into Food and Status: A Case Study from PreContact Hawaii. World Archaeology 34:484–497. Klepinger, Linda L. 1984 Nutritional Assessment from Bone. Annual Review of Anthropology 13(1):75–96. Klingelhofer, Eric 2008 Geophysics and the Search for Raleigh’s Outpost on Trinidad. In Archaeology and Geoinformatics: Case Studies from the Caribbean, edited by Basil A. Reid, pp. 155–169. The University of Alabama Press, Tuscaloosa. Kolb, Michael J., and James E. Snead 1997 It’s a Small World after All: Comparative Analyses of Community Organization in Archaeology. American Antiquity 62(4):609–628. Krause, Richard A. 1989 Coffee, Sugar and Baked Clay: From Prehistory to History in Puerto Rico’s Cerrillos River Valley. OSM Archaeological Consultants, Moundville, Alabama. Submitted to the U.S. Army Corps of Engineers, Jacksonville District, Jacksonville, Florida. Krogman, Wilton M., and M. Yasar Iscan 1986 The Human Skeleton in Forensic Medicine. Charles C. Thomas, Springfield, Illinois. Krueger, Harold W., and Charles H. Sullivan 1984 Models for Carbon Isotope Fractionation between Diet and Bone. In Stable Isotopes in Nutrition, edited by J. Turnlund and P. E. Johnson, pp. 205–222. ACS Symposium Series Vol. 258. American Chemical Society, Washing ton, D.C. Krushensky, Richard D., and Watson H. Monroe 1975 Geological Map of the Ponce Quadrangle, Puerto Rico. U.S. Geological Survey, Reston, Virginia. Kvamme, Kenneth 2001 Current Practices in Archaeogeophysics: Magnetics, Resistivity, Conductivity, and Ground-Penetrating Radar. In Earth Sciences and Archaeology, edited by Paul Goldberg, Vance T. Holliday, and C. Reid Ferring, pp. 353–384. Kluwer Academic/Plenum, New York. Laffoon, Jason, Sloan W. Williams, William J. Pestle, and L. Antonio Curet 2006 Preliminary Analysis of Ancient DNA from Prehistoric Populations of Puerto Rico. Poster presented at the Annual Meeting of the Society for American Archaeology, San Juan, Puerto Rico. Larsen, Clark S. 1997 Bioarchaeology: Interpreting Behavior from the Human Skeleton. Cambridge University Press, Cambridge. Larsen, Matthew C., and Angel J. Torres-Sanchez 1998 The Frequency and Distribution of Recent Landslides in Three Montane Tropical Regions of Puerto Rico. Geomorphology 24:309–331. Las Casas, Bartolomé 1967 Apologética historia sumaria. Universidad Autónoma de México, México, D.F. 1971 History of the Indies. Translated and edited by Andree Collard. Harper and [1527– Row, New York. 1565]

304 / References Cited Layfield, J. 1995 Relación del viaje a Puerto Rico de la expedición de Sir George Clifford, [1598] tercer conde de Cumberland, escrita por el Reverendo Doctor John Layfield, capellán de la expedición. In Cronicas de Puerto Rico: Desde la conquista hasta nuestros días (1493–1955), compiled and edited by E. Fernandez Mendez, pp. 135–156. Editorial Universitaria, San Juan, Puerto Rico. Lazzari, Eugene P. 1976 Dental Biochemistry. Lea and Febiger, Philadelphia. Liogier, Henri Alain 1988 Descriptive Flora of Puerto Rico and Adjacent Islands: Spermatophyta: Vol. 2. Leguminosae to Anacardiaceae. Editorial de la Universidad de Puerto Rico, Río Piedras. Liogier, Henri A., and Luis F. Martorell 2000 Flora of Puerto Rico and Adjacent Islands: A Systematic Synopsis. 2nd ed. Editorial de la Universidad de Puerto Rico, Río Piedras. Little, Elbert L., Jr., and Frank H. Wadsworth 1964 Common Trees of Puerto Rico and the Virgin Islands. U.S. Department of Agriculture, Forest Service, Agriculture Handbook No. 249. Government Printing Office, Washington, D.C. Little, Elbert L., Jr., Frank H. Wadsworth, and Jose Marrero 2001 Arboles comunes de Puerto Rico y las Islas Vírgenes. 2nd ed. Editorial de la Universidad de Puerto Rico, San Juan. Little, Elbert L., Jr., Roy O. Woodbury, and Frank H. Wadsworth 1974 Trees of Puerto Rico and the Virgin Islands, Vol. 2. U.S. Department of Agriculture, Forest Service, Agriculture Handbook No. 449. Government Printing Office, Washington, D.C. Longwood, Franklin R. 1962 Present and Potential Commercial Timbers of the Caribbean, with Special Reference to the West Indies, the Guianas, and British Honduras. U.S. Department of Agriculture, Forest Service, Agriculture Handbook No. 207. Government Printing Office, Washington, D.C. López Sotomayor, Diana 1975 Vieques: Un momento en su historia. Unpublished Master’s thesis, Escuela Nacional de Antropología e Historia, Mexico City. Lovén, Sven 1935 The Origins of Tainan Culture, West Indies. Elanders Bokfryckeri Akfiebolag, Göteborg, Sweden. Lugo, A. E. 2005 Los bosques. In Biodiversidad de Puerto Rico: Vertebrados terrestres y ecosistemas, edited by R. L. Joglar, pp. 395–544. Serie de Historia Natural, Instituto de Cultura Puertorriqueña, San Juan. Luna Calderón, Fernando 1976 Atlas de paleopatología ósea. Serie Científica No. 2. Editorial de la Universidad Central del Este, San Pedro de Macorís, Dominican Republic. 1988 Enfermedades en las osamentas indígenas de la isla de Santo Domingo. Boletín del Museo del Hombre Dominicano 15(21):79–83. 1993 Paleopatología de los grupos Taínos de la Española. In La cultura taína, pp. 165–174. Sociedad Estatal Quinto Centenario, Turner Libros, Madrid.

References Cited / 305 2001

Antropología y paleopatología en grupos pre-agrícolas y agrícolas. In Culturas aborígenes del Caribe, pp. 191–198. Federación Internacional de Sociedades Científicas and Banco Central de la República Dominicana, Santo Domingo. Lundberg, Emily R. 1985 Settlement Pattern Analysis for South-Central Puerto Rico (Appendix L). In Archaeological Data Recovery at El Bronce, Puerto Rico. Final Report, Phase 2, edited by L. S. Robinson, E. R. Lundberg, and J. B. Walker, pp. L1–L32. Submitted to the U.S. Army Corps of Engineers, Jacksonville District, Jacksonville, Florida. McClung, B. G., and E. Hughes 1985 Human Osteological Remains from El Bronce Archaeological Site, Puerto Rico (Appendix K). In Archaeological Data Recovery at El Bronce, Puerto Rico. Final Report, Phase 2, edited by L. S. Robinson, E. R. Lundberg, and J. B. Walker, pp. K1–K27. Submitted to the U.S. Army Corps of Engineers, Jacksonville District, Jacksonville, Florida. McGuire, Randall H., and Dean J. Saitta 1996 Although They Have Petty Captains, They Obey Them Badly: The Dialectics of Prehispanic Western Pueblo Social Organization. American Antiquity 61:197–216. Maiz López, Edgar 2002 El sitio arqueológico Hernández Colón de Puerto Rico: Actividades subsistenciales de los antiguos habitantes del valle del Río Cerrillos-Bucaná, Ponce Puerto Rico. Unpublished Master’s thesis, Centro de Estudios Avanzados de Puerto Rico y el Caribe, San Juan, Puerto Rico. Maiz López, Edgar, and Eduardo Questell Rodríguez 1990 Reconocimiento arquelogico preliminar de la cuenca hidrográfica del Rio Yauco. In Proceedings of the Eleventh International Congress for Caribbean Archaeology, edited by A. G. Pantel Tekakis, I. Vargas Arenas, and M. Sanoja Obediente, pp. 312–327. La Fundación Arqueológica, Antropológica e Histórica de Puerto Rico, San Juan. Mallory-Greenough, Leanne M., John D. Greenough, and J. Victor Owen 2000 The Origin and Use of Basalt in Old Kingdom Funerary Temples. Geoarchaeology 15:315–330. Mann, Robert W., and Sean P. Murphy 1990 Regional Atlas of Bone Disease: A Guide to Pathologic and Normal Variation in the Human Skeleton. Charles C. Thomas, Springfield, Illinois. March, James G., and Catherine M. Pringle 2003 Food Web Structure and Basal Resource Utilization along a Tropical Island Stream Continuum, Puerto Rico. Biotropica 35:84–93. Marino, Bruno D., and Michael B. McElroy 1991 Isotopic Composition of Atmospheric CO2 Inferred from Carbon in C4 Plant Cellulose. Nature 349:127–131. Mariotti, André 1983 Atmospheric Nitrogen Is a Reliable Standard for Natural 15N Abundance Measurements. Nature (London) 303:685–687. Marquardt, William H. 2001 The Emergence and Demise of the Calusa. In Societies in Eclipse: Archae-

306 / References Cited ology of the Eastern Woodlands Indians, A.D. 1400–1700, edited by D. S. Brose, C. W. Cowan, and R. C. Mainfort Jr., pp. 157–171. Smithsonian Institution Press, Washington, D.C. 2004 Calusa. In Southeast, edited by R. D. Fogelson, pp. 204–212. Handbook of North American Indians, Vol. 14, William Sturtevant, general editor. Smithsonian Institution Press, Washington, D.C. Márquez, M., and P. Hernández 2001 Principios básicos teóricos y metodológicos de la paleodemografía. Instituto Nacional de Antropología e Historia, Mexico City. Mason, J. Alden 1941 A Large Archaeological Site at Capá, Utuado, with Notes on Other Porto Rican Sites Visited in 1914–1915. Scientific Survey of Porto Rico and the Virgin Islands, Vol. 18, pt. 2. New York Academy of Science, New York. Miller, Suzanne, Fiona M. McGibbon, David H. Caldwell, and Nigel A. Ruckley 2006 Geological Tools to Interpret Scottish Medieval Carved Sculpture: Combined Petrological and Magnetic Susceptibility Analysis. Geological Society Special Publications 257:283–305. Mintz, Sidney W., and Christine M. Du Bois 2002 The Anthropology of Food and Eating. Annual Review of Anthropology 31(1):99–119. Miriello, Domenico, and Gino Mirocle Crisci 2006 The Stones of the Svevian Castle of Rocca Imperiale (Calabria, Southern Italy): Characterization and Provenance. Periodico di Mineralogia 75:205– 215. Moerman, Daniel E. 1986 Medicinal Plants of Native America, Vol. 1. Technical Reports No. 19. University of Michigan Museum of Anthropology, Ann Arbor. Molina, Daniel, and Diana López 1995 Proyecto de supervisión arqueológica (Monitoreo) en el sitio arqueológico Caracoles, Plaza Caribe, Ponce, Puerto Rico, Vols. 1 and 2. Submitted to Consejo para la Protección del Patrimonio Arqueológico Terrestre de Puerto Rico and Programa de Arqueología del Instituto de Cultura Puertorriqueña, San Juan. Morbán, L. F., A. R. Kasser, and A. E. Kasser 1976 Enfermedades de los niños aborígenes de la isla de Santo Domingo. Boletín del Museo del Hombre Dominicano 4(8):163–170. Morgan, Gary S., and Charles A. Woods 1986 Extinction and the Zoogeography of West Indian Land Mammals. Biological Journal of the Linnean Society 28:167–203. Moscoso, Francisco 1986 Tribu y clases en el Caribe antiguo. Universidad Central del Este, San Pedro de Macorís, Dominican Republic. 1999 Arcaicos de Angostura, pasado remoto de Puerto Rico: Diálogo de Francisco Moscoso, historiador, con Carlos M. Ayes Suárez y Ovidio Dávila, Arqueólogos. Sociedad de Investigaciones Arqueológicas e Históricas Seboruco, Vega Baja, Puerto Rico. Munizaga, J. R. 1992 Antropología física de los Andes del Sur. In Prehistoria Suramericana: Nuevas perspectivas, edited by B. J. Meggers, pp. 65–75. Taraxacum, Washington, D.C.

References Cited / 307 Murdock, George P. 1949 Social Structure. 1st ed. MacMillan, New York. Murphy, Peter G., and Ariel E. Lugo 1986 Ecology of Tropical Dry Forest. Annual Review of Ecology and Systematics 17:67–88. Murphy, Peter G., Ariel E. Lugo, A. J. Murphy, and D. C. Nepstad 1995 The Dry Forests of Puerto Rico’s South Coast. In Tropical Forests: Management and Ecology, edited by A. E. Lugo and C. Lowe, pp. 178–209. SpringerVerlag, New York. Narganes Storde, Yvonne M. 1985 Restos faunísticos vertebrados de Sorcé, Vieques, Puerto Rico. In Proceedings of the Tenth International Congress for the Study of the Pre-Columbian Cultures of the Lesser Antilles, edited by L. Allaire and F. M. Mayer, pp. 251–264. Université de Montréal, Montréal. 1993a Fauna y cultura indígena de Puerto Rico. Museo de Historia, Antropología y Arte, Universidad de Puerto Rico, Recinto de Río Piedras. 1993b Los restos faunisticos de sitio de Puerto Ferro, Vieques, Puerto Rico. In Proceedings of the Fourteenth International Congress for Caribbean Archaeology, edited by A. Cummins and P. King, pp. 94–114. International Association for Caribbean Archaeology, Barbados. National Academy of Sciences, Panel on Firewood Crops 1980 Firewood Crops: Shrub and Tree Species for Energy Production. National Academy of Sciences, Washington, D.C. 1983 Firewood Crops: Shrub and Tree Species for Energy Production, Vol. 2. National Academy of Sciences, Washington, D.C. Native American Ethnobotany Database ca. Native American Ethnobotany: A Database of Plants Used as Drugs, Foods, 1990– Dyes, Fibers, and More, by Native Peoples of North America. Electronic 2009 document, http://herb.umd.umich.edu/, accessed July 2007. Newsom, Lee A. 1987 Analysis of Wood Charcoal from Site PO-21, Cerrillos River, Puerto Rico. In Data Recovery Excavations at Site PO-21, Cerrillos River Valley, Puerto Rico, Appendix C. Garrow and Associates, Inc., Atlanta. 1988 Archaeobotanical Analysis of Three Features from El Fresal, Cuyon, Aibonito. In Mitigacion arqueologica franja del yacimiento Area B, Barrio Cuyon, Aibonito, Puerto Rico, edited by M. Meléndez Maíz, pp. C1–C13. Report submitted to the Departmento de Agricultura, Estado Libre Asociado, Puerto Rico. 1990 PO-39, Cerrillos River Archaeobotanicals Sample Assemblage. Unpublished laboratory data, 18 April 1990. Department of Anthropology, Pennsylvania State University, University Park. 1992 Archaeobotanical Analysis of Flotation Samples from Site PO-38, Cerrillos River Valley, Puerto Rico. In Phase II Archaeological Recovery at PO-38, El Parking Site, Barrio Maraguez, Ponce, Puerto Rico, pp. 318–330. Garrow and Associates, Inc., Atlanta. 1993 Native West Indian Plant Use. Ph.D. dissertation, University of Florida, Gainesville. University Microfilms, Ann Arbor. 1999 Mangroves and Root Crops: The Archaeobotanical Record from En Bas Saline, Haiti. In Proceedings of the Sixteenth Congress of the International Association for Caribbean Archaeology, edited by R. Gerard, pp. 52–66. Guadeloupe.

308 / References Cited 2006

Caribbean Maize: First Farmers to Columbus. In Histories of Maize, edited by R. Blake, J. Staller, and R. Tikot, pp. 325–335. Academic Press, New York. 2007 Tibes Paleoethnobotany: Report on Investigations 1995–2003. Report on file with the author, Department of Anthropology, Pennsylvania State University, University Park, and with L. A. Curet, Department of Anthropology, Field Museum of Natural History, Chicago. 2008 Caribbean Paleoethnobotany: Present Status and New Horizons (Understanding the Evolution of an Indigenous Ethnobotany). In Crossing the Borders: New Methods and Techniques in the Study of Archaeological Materials from the Caribbean, edited by C. L. Hofman, M. L. P. Hoogland, and A. L. van Gijn, pp. 173–194. The University of Alabama Press, Tuscaloosa. 2009 Paleoethnobotanical Research at Tibes, Ponce, Puerto Rico. Report submitted to the Consejo para la Protección del Patrimonio Arqueológico Terrestre de Puerto Rico, San Juan. Electronic copy available at http://www.anthro.psu .edu/faculty_staff/newsom.shtml. Newsom, Lee A., and L. Antonio Curet 2000 Report on the 1998 and 1999 Field Seasons of the Proyecto Arqueologico del Centro Ceremonial de Tibes. Project report to the Committee for Research and Exploration, The National Geographic Society, Washington, D.C. 2003a Biodiversity and Natural Capital: Toward Understanding the Ecological Economics of the Tibes Archaeological Site. Proceedings of the Twenty-Ninth International Congress for Caribbean Archaeology 9(1):156–167. Museum Arqueological Aruba, Oranjestad. 2003b Investigations of Household and Communal Deposits at Tibes, Puerto Rico (Grant #7276-02). Report submitted to the Committee for Research and Exploration, The National Geographic Society, Washington, D.C. Newsom, Lee A., and Kathleen A. Deagan 1994 Zea mays in the West Indies: The Archaeological and Early Historic Record. In Corn and Culture in the Prehistoric New World, edited by S. Johannessen and C. A. Hastorf, pp. 203–218. Westview Press, Boulder, Colorado. Newsom, Lee A., and Elizabeth S. Wing 2004 On Land and Sea: Native American Uses of Biological Resources in the West Indies. The University of Alabama Press, Tuscaloosa. Newsome, Seth D., Donald L. Phillips, Brendan J. Culleton, Tom P. Guilderson, and Paul L. Koch 2004 Dietary Reconstruction of an Early Holocene Human Population from the Central California Coast: Insights from Advanced Stable Isotope Mixing Models. Journal of Archaeological Science 31:1101–1115. Nielsen, Axel E. 1991 Trampling the Archaeological Record: An Experimental Study. American Antiquity 56(3):483–503. Nieves-Rivera, A. M., José Muñoz-Vasquez, and C. Betancourt-López 1995 Hallucinogens Used by the Taíno Indians in the West Indies. Atenea: Facultad de Artes y Ciencias (Universidad de Puerto Rico, Mayagüez) 15(1–2):125–139. Norr, Lynette 2002 Bone Isotopic Analysis and Prehistoric Diet at the Tutu Site. In The Tutu Archaeological Village Site: A Multidisciplinary Case Study in Human Adaptation, edited by Elizabeth Righter, pp. 263–273. Routledge, New York.

References Cited / 309 Nowak, Ronald M. 1999 Walker’s Mammals of the World. 6th ed. John Hopkins University Press, Baltimore. Nuñez Meléndez, Esteban 1992 Plantas medicinales de Puerto Rico: Folklore y fundamentos científicos. Editorial de la Universidad de Puerto Rico, Río Piedras. Ogburn, Dennis E. 2004 Evidence for Long-Distance Transportation of Building Stones in the Inka Empire, from Cuzco, Peru to Saraguro, Ecuador. Latin American Antiquity 15:419–439. Oliver, José R. 1998 El centro ceremonial de Caguana, Puerto Rico: Simbolismo iconográfico, cosmovisión y el poderío caciquil Taíno de Borínquen. BAR International Series 727. British Archaeological Reports, Oxford. 1999 The “La Hueca Problem” in Puerto Rico and the Caribbean: Old Problems, New Perspectives, Possible Solutions. In Archaeological Investigations on St. Martin (Lesser Antilles), edited by C. L. Hofman and M. L. P. Hoogland, pp. 253–297. Faculty of Archaeology, Leiden University, Leiden, the Netherlands. 2000 Gold Symbolism among Caribbean Chiefdoms: Of Feathers, Cibas, and Guanín Power among Taíno Elites. In Precolumbian Gold: Technology, Style, and Iconography, edited by C. McEwan, pp. 196–219. British Museum Press, London. 2005 The Proto-Taino Monumental Cemís of Caguana: A Political-Religious “Manifesto.” In Ancient Borinquen: Archaeology and Ethnohistory of Native Puerto Rico, edited by Peter E. Siegel, pp. 230–284. The University of Alabama Press, Tuscaloosa. 2009 Caciques and Cemí Idols: The Web Spun by Taíno Rulers in Hispaniola and Puerto Rico. The University of Alabama Press, Tuscaloosa. Oliver, José R., Juan A. Rivera Fontán, and Lee A. Newsom 1999 Arqueología del Barrio Caguana, Puerto Rico: Resultados preliminaries de las temporadas 1996–1997. In Segundo encuentro de investigadores: Trabajos de investigación arqueológica en Puerto Rico, edited by J. A. Rivera Fontán, pp. 7–26. Publicación Ocasional de la División de Arqueología, Instituto de Cultura Puertorriqueña, San Juan. Ortner, Donald J. 2003 Identification of Pathological Conditions in Human Skeletal Remains. 2nd ed. Academic Press, San Diego, California. Osborn, Alan J. 1977 Strandloopers, Mermaids, and Other Fairy Tales: Ecological Determinants of Marine Resource Utilization—The Peruvian Case. In For Theory Building in Archaeology, edited by Lewis Binford, pp. 157–206. Academic Press, New York. Pagán Jiménez, Jaime, Miguel Rodríguez López, Luis A. Chanlatte Baik, and Yvonne Narganes Storde 2005 La temprana introducción y uso de algunas plantas domésticas, silvestres, y cultivos en Las Antillas precolombinas. Una primera revaloración desde la perspectiva del “Arcaico” de Vieques y Puerto Rico. Diálogo Antropológico 3(10):7–33.

310 / References Cited Pané, Fray Ramón 2001 Relacion acerca de las antigüedades de los indios. Siglo Veintiuno Editores, [1505] America Nuestra, Mexico. Pantel, A. Gus 1978 Portugues and Bucana Rivers, Puerto Rico: Cultural Resources Survey. La Fundación Arqueológica, Antropológica e Histórica de Puerto Rico, La Universidad de Puerto Rico, San Juan. Submitted to the U.S. Army Corps of Engineers, Jacksonville District, Jacksonville, Florida. Parker, Patrick L. 1964 The Biogeochemistry of the Stable Isotopes of Carbon in a Marine Bay. Geochimica et Cosmochimica Acta 28:1155–1164. Pauketat, Timothy R. 2000 Practice and History in Archaeology. Anthropological Theory 1:73–98. 2007 Chiefdoms and Other Archaeological Delusions. Altamira Press, Lanham, Maryland. Pearsall, Deborah M. 1985 Analysis of Soil Phytoliths and Botanical Macroremains from El Bronce Archaeological Site, Ponce, Puerto Rico (Appendix B). In Archaeological Data Recovery at El Bronce, Puerto Rico. Final Report, Phase 2, edited by L. S. Robinson, E. R. Lundberg, and J. B. Walker, pp. B1–B46. Submitted to the U.S. Army Corps of Engineers, Jacksonville District, Jacksonville, Florida. 2000 Paleoethnobotany: A Handbook of Procedures. 2nd ed. Academic Press, San Diego, California. 2002 Analysis of Charred Botanical Remains from the Tutu Site, U.S. Virgin Islands. In The Tutu Archaeological Village Site: A Multidisciplinary Case Study in Human Adaptation, edited by Elizabeth Righter, pp. 109–134. Routledge, New York. Pessagno, Emile A., Jr. 1960 Geology of Ponce-Coamo Area, Puerto Rico. Unpublished Ph.D. dissertation, Department of Geology, Princeton University, Princeton, New Jersey. Phillips, Donald L., and Jillian W. Gregg 2003 Source Partitioning Using Stable Isotopes: Coping with Too Many Sources. Oecologia 136:261–269. Phillips, Donald L., Seth D. Newsome, and Jillian W. Gregg 2005 Combining Sources in Stable Isotope Mixing Models: Alternative Methods. Oecologia 144(4):520–527. Picó, Rafael 1950 The Geographic Regions of Puerto Rico. Universidad de Puerto Rico, Río Piedras. 1974 Geography of Puerto Rico. Aldine, Chicago. Pina, Plinio P. 1972 Las deformaciones intencionales del cuerpo humano en las Antillas. Boletín del Museo del Hombre Dominicano 1:9–19. Preston, Jeremy, Ian Meighan, Derek Simpson, and Malcolm Hole 2002 Mineral Chemical Provenance of Neolithic Pitchstone Artefacts from Ballygalley, County Antrim, Northern Ireland. Geoarchaeology 17:219–236. Quitmyer, Irvy R. 1985 Zooarchaeological Methods for the Analysis of Shell Middens at Kings Bay.

References Cited / 311 In Aboriginal Subsistence and Settlement Archaeology of the Kings Bay Locality, edited by W. H. Adams, pp. 33–48. Report of Investigations, Vol. 2. Department of Anthropology, University of Florida, Gainesville. Quitmyer, Irvy R., and Laura Kozuch 1996 Phase II Zooarchaeology at Finca Valencia (NCS-1) and Site NCS-4, Northwest Puerto Rico (with a Contribution from Elizabeth S. Wing). Report submitted to LAW Caribe Environmental and Engineering, Inc., San Juan, Puerto Rico, and on file at the Environmental Archaeology Laboratory, Florida Museum of Natural History, Gainesville. Raffaele, Herbert 1989 A Guide to the Birds of Puerto Rico and the Virgin Islands. Rev. ed. Princeton University Press, Princeton, New Jersey. Rainey, Froelich G. 1935 Puerto Rican Archaeology. Yale University, New Haven, Connecticut. 1940 Porto Rican Archaeology. Scientific Survey of Porto Rico and the Virgin Islands, Vol. 18, pt. 1. New York Academy of Science, New York. Record, Samuel J., and Robert W. Hess 1986 Timbers of the New World. Reprinted, Ayer Company, Salem, New Hampshire. Originally published 1943, Yale University Press, New Haven, Connecticut. Redmond, Elsa M., and Charles S. Spencer 1994 The Cacicazgo: An Indigenous Design. In Caciques and Their People: A Volume in Honor of Ronald Spores, edited by J. Marcus and J. F. Zeitlin, pp. 189– 225. Anthropological Papers No. 89. University of Michigan Museum of Anthropology, Ann Arbor. 2007 Archaeological Survey in the High Llanos and Andean Piedmont of Barinas, Venezuela. Anthropological Papers No. 86. American Museum of Natural History, New York. Rehm, Sigmund, and Gustav Espig 1991 The Cultivated Plants of the Tropics and Subtropics: Cultivation, Economic Value, and Utilization. Institute of Agronomy in the Tropics, University of Göttingen. Verlag Josef Margraf, Weikersheim, West Germany. Rice-Snow, Scott, Melissa J. Castor, Andrew K. Castor, Jeffry D. Grigsby, Richard H. Fluegeman, and L. Antonio Curet 2006 Boulder Lithology Survey, Tibes Civic-Ceremonial Site, Puerto Rico. Paper presented at the 71st Annual Meeting of the Society for American Archaeology, San Juan. Rich, Fred 1999 Report on the Contents and Characteristics of Pollen Samples from the Tibes Archaeological Site. Report submitted to L. A. Newsom and L. A. Curet, Proyecto Arqueológico del Centro Ceremonial de Tibes; on file at the Departments of Anthropology, Pennsylvania State University, University Park, and Field Museum of Natural History, Chicago. Righter, Elizabeth (editor) 2002 The Tutu Archaeological Village Site: A Multidisciplinary Case Study in Human Adaptation. Routledge, New York. Rivera Fontán, Juan A. 1992 Proyecto arqueológico Caguana 92: Reconocimiento sistemático de los recur-

312 / References Cited sos arqueológicos del Parque Ceremonial de Caguana, Utuado, P.R. División de Arqueología, Instituto de Cultura Puertorriqueña, San Juan. 1999 Los bateyes de Caguana. Revista Cultura 6. Instituto de Cultura Puertorriqueña, San Juan. Rivera Fontán, Juan A., and José R. Oliver 2005 Impactos y patrones de ocupación histórica jíbara sobre componentes taínos: El sitio “Vega de Nelo Vargas” (UTU-27), Barrio Caguana, Municipio de Utuado, Puerto Rico. In Proceedings of the Twentieth International Congress for Caribbean Archaeology, edited by Glenis Tavares María and Manuel García Arévalo, pp. 1–14. Museo del Hombre Dominicano and Fundación García Arévalo, Santo Domingo. Rivera Fontán, Juan A., and Daniel S. Silva 1997 Proyecto arqueológico Bo. Quemado, Mayagüez: Bate Delfín del Yagüez. In Ocho trabajos de investigación arqueológica en Puerto Rico: Segundo encuentro de investigadores, edited by J. A. Rivera Fontán, pp. 53–64. Publicación Ocasional de la División de Arqueología, Instituto de Cultura Puertorriqueña, San Juan. 2002 Batey Delfín del Yagüez: Un asentamiento taíno en el cauce medio del Río Yagüez. In Cuarto encuentro de investigadores: Trabajos de investigación arqueológica, edited by C. A. Pérez Merced and J. A. Rivera Fontán, pp. 69– 84. Publicación Ocasional de la División de Arqueología, Instituto de Cultura Puertorriqueña, San Juan. Rivero de la Calle, Miguel 1949 Deformación craneana de los aborígenes de Cuba: Estudio comparativo. Unpublished Ph.D. dissertation, Universidad de la Habana, Cuba. 1985 Nociones de anatomia humana aplicadas a la Arqueología. Editorial Científico-Técnica, Ciudad de la Habana, Cuba. 1987 Paleopatología de los aborígenes de Cuba. In Estudios de antropología biológica, edited by M. E. Saenz and X. Lizarraga, pp. 475–498. Universidad Nacional Autónoma de México, México City. 1993 Antropología física de los Taínos. In La cultura taína, pp. 153–161. Sociedad Estatal Quinto Centenario, Turner Libros, Madrid. Roberts, Charlotte, and Keith Manchester 1999 The Archaeology of Disease. 2nd ed. Cornell University Press, Ithaca, New York. Robinson, Linda S., Emily R. Lundberg, and Jeffery B. Walker 1985 Archaeological Data Recovery at El Bronce, Puerto Rico. Final Report, Phase 2. Submitted to the U.S. Army Corps of Engineers, Jacksonville District, Jacksonville, Florida. Rodríguez, Miguel A. 1983 Prehistoria de Collores. Unpublished Master’s thesis, Centro de Estudios Avanzados de Puerto Rico y el Caribe, San Juan, Puerto Rico. 1985 Cultural Resources Survey at Camp Santiago, Salinas, Puerto Rico. In Proceedings of the Tenth International Congress for the Study of the PreColumbian Cultures of the Lesser Antilles, edited by L. Allaire and F. M. Mayer, pp. 23–44. Université de Montréal, Montréal. 1989 The Zoned Incised Crosshatch (ZIC) Ware of Early Precolumbian Ceramic Age Sites in Puerto Rico and Vieques Island. In Early Ceramic Population Lifeways and Adaptive Strategies in the Caribbean, edited by P. E. Siegel,

References Cited / 313 pp. 249–266. BAR International Series 506. British Archaeological Reports, Oxford. 1991 Arqueología de Punta Candelero. In Proceedings of the Thirteenth International Congress for Caribbean Archaeology, edited by E. N. Ayubi and J. B. Haviser, pp. 605–627. Reports of the Archaeological-Anthropological Institute of the Netherlands Antilles No. 9. Curaçao. 1992 Diversidad cultural en la tardía prehistoria del este de Puerto Rico. Revista del Centro de Estudios Avanzados de Puerto Rico y el Caribe 15:58–74. 1997 Maruca, Ponce. In Ocho trabajos de investigación arqueológica en Puerto Rico: Segundo encuentro de investigadores, edited by J. A. Rivera Fontán, pp. 17–30. Publicación Ocasional de la División de Arqueología, Instituto de Cultura Puertorriqueña, San Juan. 1999 Excavations at Maruca, a Preceramic Site in Southern Puerto Rico. In Proceedings of the Seventeenth Congress of the International Association for Caribbean Archaeology, edited by J. H. Winter, pp. 166–180. Molloy College, Rockville Centre, New York. Rodríguez, S. R., and M. A. Terrazas 2003 Presencia de huellas de corte como evidencia de tratamiento mortuorio en un cráneo agroalfarero del oriente de Cuba. In Estudios de antropología biológica, Vol. 11, Pt. 2, edited by C. Serrano, S. López, and F. Pedraza, pp. 1029– 1035. Universidad Nacional Autónoma de México, Instituto Nacional de Antropología e Historia, and Asociación Mexicana de Antropología Biológica, México City. Rodríguez Gracia, Luis A. 1998 La cerámica de Tibes: Descripción y uso. Manuscript on file, Centro Ceremonial Indígena de Tibes, Ponce, Puerto Rico, and with A. Curet, Field Museum of Natural History, Chicago. Rodríguez Meléndez, Yasha N. 2007 Social Life of Bateyes: Archaeology, Preservation and Heritage in Puerto Rico. Unpublished Ph.D. dissertation, Department of Anthropology, Cornell University, Ithaca, New York. Rodríguez Ramos, Reniel 2001 Lithic Reduction Trajectories at La Hueca and Punta Candelero Sites (Puerto Rico). Unpublished Master’s thesis, Department of Anthropology, Texas A&M University, College Station. 2002a Dinámicas de intercambio en el Puerto Rico prehispánico. El Caribe Arqueológico 6:16–22. 2002b Una perspectiva diacrónica de la explotación del pedernal en Puerto Rico. Boletín del Museo del Hombre Dominicano 32:167–192. 2005 The Crab-Shell Dichotomy Revisited: The Lithics Speak Out. In Ancient Borinquen: Archaeology and Ethnohistory of Native Puerto Rico, edited by Peter E. Siegel, pp. 1–54. The University of Alabama Press, Tuscaloosa. 2007 Puerto Rican Precolonial History Etched in Stone. Unpublished Ph.D. dissertation, Department of Anthropology, University of Florida, Gainesville. 2008 From the Guanahatabey to the Archaic of Puerto Rico: The Non-Evident Evidence. Ethnohistory 55(3):393–415. Rodríguez Ramos, Reniel, and Jaime Pagán Jiménez 2006 Interacciones multivectoriales en el Circum-Caribe precolonial: Un vistazo desde las Antillas. Caribbean Studies 34(2):103–143.

314 / References Cited 2007

Las Antillas en el contexto del Circum-Caribe: Cincuenta años después. In Proceedings of the Twenty-First International Congress for Caribbean Archaeology, edited by B. Reid, H. Petitjean Roget, and A. Curet, pp. 778–786. University of the West Indies, St. Augustine, Trinidad. Rodríguez Ramos, Reniel, Josh M. Torres, and José R. Oliver 2010 Rethinking Time in Caribbean Archaeology. In The Prehistoric Settlement of the Caribbean, edited by S. Fitzpatrick and A. H. Ross. University Press of Florida, Gainesville. Roe, Peter G. 1989 A Grammatical Analysis of Cedrosan Saladoid Vessel Form Categories and Surface Decoration: Aesthetic and Technical Styles in Early Antillean Ceramics. In Early Ceramic Population Lifeways and Adaptive Strategies in the Caribbean, edited by P. E. Siegel, pp. 267–382. BAR International Series 506. British Archaeological Reports, Oxford. Rofes, Juan 2000 Sacrificio de Cuyes en El Yaral, comunidad prehispánica del extremo sur peruano. Bulletin de l’Institut Français d’Etudes Andines 29(1):1–12. Roosevelt, Anna C. 1980 Parmana: Prehistoric Maize and Manioc Subsistence along the Amazon and Orinoco. Academic Press, New York. 1990 El método y la teoría de la arqueología de las tierras bajas tropicales. In Proceedings of the Eleventh International Congress for Caribbean Archaeology, edited by A. G. Pantel Tekakis, I. Vargas Arenas, and M. Sanoja Obediente, pp. 27–40. La Fundación Arqueológica, Antropológica e Histórica de Puerto Rico, San Juan. Ross, Eric B. 1987 An Overview of Trends in Dietary Variation from Hunter-Gatherer to Modern Capitalist Societies. In Food and Evolution: Toward a Theory of Human Food Habits, edited by Marvin Harris and Eric B. Ross, pp. 7–55. Temple University Press, Philadelphia. Roth, Walter Edmund 1924 An Introductory Study of the Arts, Crafts, and Customs of the Guiana Indians. Annual Report of the Bureau of American Ethnology 38. Smithsonian Institution, Washington, D.C. Rountree, Helen C., and E. Randolph Turner 1998 The Evolution of the Powhatan Paramount Chiefdom in Virginia. In Chiefdoms and Chieftaincy in the Americas, edited by E. M. Redmond, pp. 265–296. University Press of Florida, Gainesville. Rouse, Irving 1948 The Arawak. In The Circum-Caribbean Tribes, edited by J. Steward, pp. 50– 46. Handbook of South American Indians, Vol. 4, Bulletin of the Bureau of American Ethnology No. 143. Smithsonian Institution, Washington, D.C. 1952 Porto Rican Prehistory. Scientific Survey of Porto Rico and the Virgin Islands, Vol. 18, pts. 3–4. New York Academy of Science, New York. 1964 Prehistory of the West Indies. Science 144:499–513. 1982 Ceramic and Religious Development in the Greater Antilles. Journal of New World Archaeology 5:45–55. 1986 Migrations in Prehistory: Inferring Population Movement from Cultural Remains. Yale University Press, New Haven, Connecticut.

References Cited / 315 1992

The Tainos: Rise and Decline of the People Who Greeted Columbus. Yale University Press, New Haven, Connecticut. Rouse, Irving, and Ricardo E. Alegría 1990 Excavations at María de La Cruz Cave and Hacienda Grande Village Site, Lozía, Puerto Rico. Yale University Publications in Anthropology No. 80. Department of Anthropology, Yale University, New Haven, Connecticut. Russ, John C. 1995 The Image Processing Handbook. 2nd ed. CRC Press, Boca Raton, Florida. Sage, Rowan F., Meirong Li, and Russell K. Monson 1999 The Taxonomic Distribution of C4 Photosynthesis. In C4 Plant Biology, edited by Rowan F. Sage and Russell K. Monson, pp. 551–584. Academic Press, San Diego, California. Saitta, Dean J. 1997 Power, Labor, and the Dynamics of Change in Chacoan Political Economy. American Antiquity 62:7–26. Sandweiss, Daniel S., and Elizabeth S. Wing 1997 Ritual Rodents: The Guinea Pigs of Chincha, Peru. Journal of Field Archaeology 24:47–58. Santos, A. L., P. Allsworth-Jones, and E. Rodrigues 2002 Pathological Evidence in Pre-columbian Human Remains from the Lee Collection (Jamaica). Antropologia Portuguesa 19:121–138. Sassaman, Kenneth E. 2005 Poverty Point as Structure, Event, Process. Journal of Archaeological Method and Theory 12(4):335–364. Sassaman, Kenneth E., and A. Randall 2007 The Alterity of Complexity and the Illusion of Change. Paper presented at the 72nd Annual Meeting of the Society for American Archaeology, Austin, Texas. Sauer, Carl O. 1966 The Early Spanish Main. University of California Press, Berkeley. Scheuer, Louise, and Sue Black 2000 Developmental Juvenile Osteology. Elsevier Academic, San Diego, California. Schiffer, Michael B. 1987 Formation Processes of the Archaeological Record. University of New Mexico Press, Albuquerque. Schmidt, K. P. 1928 Amphibians and Land Reptiles of Porto Rico. Scientific Survey of Porto Rico and the Virgin Islands, Vol. 10, pt. 1, pp. 3–160. New York Academy of Science, New York. Schoeller, Dale A. 1999 Isotope Fractionation: Why Aren’t We What We Eat? Journal of Archaeological Science 26(6):667–673. Schoeninger, Margaret J. 1985 Trophic Level Effects of 15N/14N and 13C/12C Ratios in Bone Collagen and Strontium Levels in Bone Mineral. Journal of Human Evolution 14:515–525. 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(4):625–639.

316 / References Cited Schoeninger, Margaret J., Michael J. DeNiro, and H. Tauber 1983 Stable Nitrogen Isotope Ratios of Bone Collagen Reflect Marine and Terrestrial Components of Prehistoric Human Diet. Science 220:1381–1383. Schultes, Richard E., Albert Hofmann, and Christian Rätsh 1998 Plants of the Gods: Their Sacred, Healing, and Hallucinogenic Powers. Healing Arts Press, Rochester, Vermont. Schultes, Richard E., and Robert F. Raffauf 1990 The Healing Forest: Medicinal and Toxic Plants of the Northwest Amazonia. Dioscorides Press, Portland, Oregon. Schwarcz, Henry P. 1991 Some Theoretical Aspects of Isotope Paleodiet Studies. Journal of Archaeological Science 18:261–275. Schwarcz, Henry P., and Margaret J. Schoeninger 1991 Stable Isotope Analyses in Human Nutritional Ecology. Yearbook of Physical Anthropology 34:283–321. Schwartz, Jeffrey H. 1995 Skeleton Keys: An Introduction to Human Skeletal Morphology, Development, and Analysis. Oxford University Press, New York. Scott, G. Richard, and Christy G. Turner 1997 The Anthropology of Modern Human Teeth. Cambridge University Press, Cambridge. Scudder, Sylvia J. 1999 Soils Investigations at the Tibes Site, Ponce, Puerto Rico. Report submitted to L. A. Newsom and L. A. Curet, Proyecto Arqueológico del Centro Ceremonial de Tibes; on file at the Departments of Anthropology, Pennsylvania State University, University Park, and Field Museum of Natural History, Chicago. 2001 Soil Resources and Anthropogenic Changes at the Tibes Site, Ponce, Puerto Rico. Caribbean Journal of Science 37:30–40. Siegel, Peter E. 1989 Site Structure, Demography and Social Complexity in the Early Ceramic Age of the Caribbean. In Early Ceramic Population Lifeways and Adaptive Strategies in the Caribbean, edited by P. E. Siegel, pp. 193–245. BAR International Series 506. British Archaeological Reports, Oxford. 1991 Political Evolution in the Caribbean. In Proceedings of the Thirteenth International Congress for Caribbean Archaeology, edited by E. N. Ayubi and J. B. Haviser, pp. 232–250. Reports of the Archaeological-Anthropological Institute of the Netherlands Antilles No. 9. Curaçao. 1992 Ideology, Power, and Social Complexity in Prehistoric Puerto Rico. Ph.D. dissertation, State University of New York, Binghamton. University Microfilms International, Ann Arbor. 1996 Ideology and Culture Change in Prehistoric Puerto Rico: A View from the Community. Journal of Field Archaeology 23:313–333. 1999 Contested Places and Places of Contest: The Evolution of Social Power and Ceremonial Space in Prehistoric Puerto Rico. Latin American Antiquity 10:209–238. 2004 What Happened after AD 600 on Puerto Rico? Corporate Groups, Population Restructuring, and Post-Saladoid Changes. In Late Ceramic Age Societies

References Cited / 317 in the Eastern Caribbean, edited by A. Delpuech and C. Hoffman, pp. 87–100. BAR International Series 1273. Archaeopress, Oxford. 2005 Multiple Visions of an Island’s Past and Some Thoughts for Future Directions in Puerto Rican Prehistory. In Ancient Borinquen: Archaeology and Ethnohistory of Native Puerto Rico, edited by Peter E. Siegel, pp. 353–363. The University of Alabama Press, Tuscaloosa. Smith, Adam T. 2003 The Political Landscape. University of California Press, Berkeley. Smith, Bruce N., and Samuel Epstein 1971 Two Categories of 13C/12C Ratios for Higher Plants. Plant Physiology 47(3):380. Smith, R. C. 1990 Ships of Discovery in the Caribbean. In Proceedings of the Eleventh International Congress for Caribbean Archaeology, edited by A. G. Pantel Tekakis, I. Vargas Arenas, and M. Sanoja Obediente, pp. 529–535. La Fundación Arqueológica, Antropológica e Histórica de Puerto Rico, San Juan. Solís Magaña, Carlos 1985 Archaeological Testing and Evaluation of Sites within the Proposed Portugues and Bucana Projects, in the Barrios of Tibes and Maraguez, Municipio de Ponce, Puerto Rico. OSM Archaeological Consultants, Northport, Alabama. Submitted to the U.S. Army Corps of Engineers, Jacksonville District, Jacksonville, Florida. Spouge, J. D. 1973 Oral Pathology. C. V. Mosby, Saint Louis. Stahl, Peter W., and J. A. Zeidler 1990 Differential Bone-Refuse Accumulation in Food-Preparation and Traffic Areas on Early Ecuadorian House Floor. Latin American Antiquity 1(2):150– 169. Stanton, G. 1969 The Relation of Diet to Salivary Calculus Formation. Journal of Periodontology 40:67–172. Steel, Gentry, and Claudia A. Bramblett 1988 The Anatomy and Biology of the Human Skeleton. Texas A&M University Press, College Station. Steinbock, R. Ted 1976 Paleopathological Diagnosis and Interpretation: Bone Disease in Ancient Human Populations. Charles C. Thomas, Springfield, Illinois. Stevens-Arroyo, Antonio M. 2006 Cave of the Jagua: The Mythological World of the Tainos. University of Scranton Press, Scranton, Pennsylvania. Stokes, Anne V. 1998 A Biogeographic Survey of Prehistoric Human Diet in the West Indies Using Stable Isotopes. Ph.D. dissertation, University of Florida, Gainesville. University Microfilms International, Ann Arbor. 2005 Ceramic-Age Dietary Patterns in Puerto Rico: Stable Isotopes and Island Biogeography. In Ancient Borinquen: Archaeology and Ethnohistory of Native Puerto Rico, edited by Peter E. Siegel, pp. 185–201. The University of Alabama Press, Tuscaloosa.

318 / References Cited Sued-Badillo, Jalil 1989 La mujer indígena y su sociedad. Editorial Cultural, San Juan, Puerto Rico. Sullivan, Charles H., and Harold W. Krueger 1981 Carbon Isotope Analysis of Separate Chemical Phases in Modern and Fossil Bone. Nature 292(5821):333–335. 1983 Carbon Isotope Ratios of Bone Apatite and Animal Diet Reconstruction. Nature 301:177–178. Tabío, Ernesto E., and Estrella Rey 1966 Prehistoria de Cuba. Academia de Ciencias de Cuba, Havana. Tavares María, Glenis 1996 Límites territoriales de los aborígenes de la isla de Haití a la llegada de los españoles. In Ponencias del primer seminario de arqueología del Caribe, edited by M. Veloz Maggiolo and A. Caba Fuentes, pp. 34–47. Museo Arqueológico Regional Altos de Chavón, Dominican Republic. Taylor, Douglas MacRae 1951 The Black Carib of British Honduras. Viking Fund Publications in Anthropology 17. Wenner-Gren Foundation for Anthropological Research, New York. Thomas, David John 1982 Order without Government: The Society of the Pemon Indians of Venezuela. Illinois Studies in Anthropology No. 13. University of Illinois Press, Urbana. Thomas, Prentice M., Jr., and Mark T. Swanson 1986 A Cultural Resources Reconnaissance and Testing Program of a 100 Acre Clay Borrow Area, Ponce, Puerto Rico. New World Research, Fort Walton Beach, Florida. Tieszen, Larry L., Thomas W. Boutton, K. G. Tesdahl, and N. A. Slade 1983 Fractionation and Turnover of Stable Carbon Isotopes in Animal Tissues: Implications for δ13C Analysis of Diet. Oecologia 57(1):32–37. Torres, Joshua M. 2001 Settlement Patterns and Political Geography of the Saladoid and Ostionoid Peoples of South Central Puerto Rico: An Exploration of Prehistoric Social Complexity at a Regional Level. Unpublished Master’s thesis, Department of Anthropology, University of Colorado, Denver. 2005 Deconstructing the Polity: Socio-Political Landscapes of the Ceramic Age Peoples of South Central Puerto Rico. In Ancient Borinquen: Archaeology and Ethnohistory of Native Puerto Rico, edited by Peter E. Siegel, pp. 202–229. The University of Alabama Press, Tuscaloosa. 2008 The Tibes Archaeological Survey Project: Location, Documentation and Preliminary Evaluation of Pre-Colonial Cultural Resources. Report submitted to La Oficina Estatal de Conservación Histórica, San Juan, Puerto Rico, Contract No. 2008-155037. Torres, Joshua M., L. Antonio Curet, Scott Rice-Snow, Andrew Castor, and Melissa Castor 2007 Flesh and Stone: Labor Estimates and the Socio-Political Implications of Plaza/Ballcourt Construction in Pre-Colonial Puerto Rico. Manuscript in possession of the author. Trigger, Bruce G. 1978 The Concept of the Community. In Time and Traditions: Essays in Archaeological Thought, pp. 115–121. Columbia University Press, New York.

References Cited / 319 Turner, Christy G., and L. M. C. Machado 1983 A New Dental Wear Pattern and Evidence for High Carbohydrate Consumption in a Brazilian Archaic Skeletal Population. American Journal of Physical Anthropology 61:125–130. Tuross, Noreen, Marilyn L. Fogel, Lee Newsom, and Glen H. Doran 1994 Subsistence in the Florida Archaic: The Stable-Isotope and Archaeobotanical Evidence from the Windover Site. American Antiquity 59(2):288–303. Ubelaker, Douglas H. 1989 Human Skeletal Remains: Excavation, Analysis, Interpretation. 2nd ed. Taraxacum, Washington, D.C. Urey, Harold C. 1947 The Thermodynamic Properties of Isotopic Substances. Journal of the Chemical Society (Resumed) 1947:562–581. U.S. Geological Survey 2001 Digital Elevation Models. U.S. Department of the Interior. Distributed by GeoCommunity, http://data.geocomm.com/. Valcárcel Rojas, Roberto 1999 Banes precolombino: Jerarquía y sociedad. El Caribe Arqueológico 3:84–89. 2002 Banes precolombino: La ocupación agricultora. Ediciones Holguín, Holguín, Cuba. van Klinken, G. J. 1991 Dating and Dietary Reconstruction by Isotopic Analysis of Amino Acids of Fossil Bone Collagen—With Special Reference to the Caribbean. Publications of the Foundation for Scientific Research in the Caribbean Region, Amsterdam. 1999 Bone Collagen Quality Indicators for Palaeodietary and Radiocarbon Measurements. Journal of Archaeological Science 26(6):687–695. van Leusen, P. Martijn 1999 Viewshed and Cost Surface Analysis Using GIS (Cartagraphic Modeling in Cell Based GIS II). In New Techniques for Old Times—CAA98—Computer Applications and Quantitative Methods in Archaeology: Proceedings of the 26th Conference, Barcelona 1998, edited by Juan A. Barceló, Ivan Briz, and Assumpció Vila, pp. 215–233. BAR International Series S757. British Archaeological Reports, Oxford. van der Merwe, Nikolaas J. 1982 Carbon Isotopes, Photosynthesis, and Archaeology. American Scientist 70:596–605. van der Merwe, Nikolaas J., Anna Curtenius Roosevelt, and J. C. Vogel 1981 Isotopic Evidence for Prehistoric Subsistence Change at Parmana, Venezuela. Nature 292(5823):536–538. Varien, Mark D. 1999 Sedentism and Mobility in a Social Landscape. University of Arizona Press, Tucson. 2002 Persistent Communities and Mobile Households. In Seeking the Center Place, edited by M. D. Varien and R. H. Wilshusen, pp. 163–185. University of Utah Press, Salt Lake City. Veloz Maggiolo, Marcio 1975 Cayo Cofresí: Un sitio precerámico de Puerto Rico. Ediciones de Taller, Santo Domingo, Dominican Republic. 1977– Medioambiente y adaptación humana en la prehistoria de Santo Domingo.

320 / References Cited 1978

2 vols. Editora de la Universidad Autónoma de Santo Domingo, Santo Domingo, Dominican Republic. 1991 Panorama histórico del Caribe precolombino. Banco Central de la República Dominicana, Santo Domingo. 1993 La isla de Santo Domingo antes de Colón. Banco Central de la República Dominicana, Santo Domingo. Vento, C. E., and R. D. González 1996 Paleopatología aborigen de Cuba. El Caribe Arqueológico 1:31–38. Versteeg, Aad H., and Stephan Rostain (editors) 1996 The Archaeology of Aruba: The Tanki Flip Site. Monographic Series No. 8, Aruba Museum of Archaeology, Oranjestad. Versteeg, Aad H., and Keesh Schinkel (editors) 1992 The Archaeology of St. Eustatius: The Golden Rock Site. St. Eustatius Historical Foundation, St. Eustatius. Vescelius, Gary S. 1977 Ballcourts and Boundaries in the Eastern Antilles. Manuscript on file, Department of Anthropology, Peabody Museum, Yale University, New Haven, Connecticut. Vogel, J. C., and Nikolaas J. van der Merwe 1977 Isotopic Evidence for Early Maize Cultivation in New York State. American Antiquity 42(2):238–242. Walker, Jeffery B. 1980 Analysis and Replication of the Lithic Artifacts from the Sugar Factory Pier Site, St. Kitts, West Indies. Unpublished Master’s thesis, Department of Anthropology, Washington State University, Pullman. 1993 Stone Collars, Elbow Stones and Three-Pointers, and the Nature of Taino Ritual and Myth. Ph.D. dissertation, Washington State University, Pullman. University Microfilms, Ann Arbor. 2005 The Paso del Indio Site, Vega Baja, Puerto Rico: A Progress Report. In Ancient Borinquen: Archaeology and Ethnohistory of Native Puerto Rico, edited by Peter E. Siegel, pp. 55–87. The University of Alabama Press, Tuscaloosa. Walker, Jeffery B., and Richard Wilk 1983 Manufacture and Use-Wear Characteristics of Ethnographic, Replicated, and Archaeological Manioc Grater Board Teeth. In Obsidiana en Mesoamerica, Colección Científica, Vol. 176, pp. 459–463. Instituto Nacional de Antropología e Historia, Mexico City. Warmke, Germaine L., and R. Tucker Abbott 1961 Caribbean Seashells: A Guide to the Marine Mollusks of Puerto Rico and Other West Indian Islands, Bermuda and the Lower Florida Keys. Livingston Publishing, Narberth, Pennsylvania. 1975 Caribbean Seashells: A Guide to the Marine Mollusks of Puerto Rico and Other West Indian Islands, Bermuda and the Lower Florida Keys. Dover Publications, New York. Watters, David R. 1994 Archaeology of Trants, Montserrat, Part 1: Field Methods and Artifact Density Distributions. Annals of Carnegie Museum 63:265–295. Watts, Brandon G., Marvin E. Bennett III, Otto C. Kopp, and Gerald L. Mattingly 2004 Geochemistry and Petrography of Basalt Grindstones from the Karak Plateau, Central Jordan. Geoarchaeology 19:47–69.

References Cited / 321 Weaver, Guy G., Patrick H. Garrow, and Jose R. Oliver 1992 Phase II Archaeological Data Recovery at PO-38, El Parking Site, Barrio Maraguez, Ponce, Puerto Rico. Garrow and Associates, Memphis. Welch, Daniel P. 2001 Sacred Space and the Development of Social Complexity at the CivicCeremonial Center of Tibes, Ponce, Puerto Rico. Unpublished Master’s thesis, Department of Anthropology, Boston University, Boston. Weymouth, John W. 1986 Geophysical Methods of Archaeological Site Surveying. In Advances in Archaeological Method and Theory, Vol. 9, edited by Michael B. Shiffer, pp. 311– 395. Academic Press, New York. Wheatley, David, and Mark Gillings 2002 Spatial Technology and Archaeology: The Archaeological Applications of GIS. Taylor and Francis, London. White, Christine D., and Henry P. Schwarcz 1989 Ancient Maya Diet: As Inferred from Isotopic and Elemental Analysis of Human Bone. Journal of Archaeological Science 16:451–474. White, T. D., and P. A. Folkens 2005 The Human Bone Manual. Elsevier Academic, San Diego, California. Wilson, Samuel M. 1990 Hispaniola: Caribbean Chiefdoms in the Age of Columbus. The University of Alabama Press, Tuscaloosa. 1991 Current Research. American Antiquity 56(1):143–160. 2006 Conclusions. In The Prehistory of Nevis, a Small Island in the Lesser Antilles, pp. 215–218. Yale University Publications in Anthropology No. 87. Peabody Museum of Natural History, Yale University, New Haven, Connecticut. 2007 The Archaeology of the Caribbean. Cambridge University Press, Cambridge. Wing, Elizabeth S. 1989 Human Exploitation of Animal Resources in the Caribbean. In Biogeography of the West Indies, edited by Charles A. Woods, 137–150. Sand Hill Crane Press, Gainesville, Florida. 1996 Guinea Pig (Cavia porcellus) Remains from Finca Valencia (NCS-1), Northwest Puerto Rico (appendix). In Phase II Zooarchaeology at Finca Valencia (NCS-1) and Site NCS-4, Northwest Puerto Rico, by Irvy R. Quitmyer and Laura Kozuch. Report submitted to LAW Caribe Environmental and Engineering, Inc., San Juan, Puerto Rico, and on file at the Environmental Archaeology Laboratory, Florida Museum of Natural History, Gainesville. 2001a Native American Use of Animals in the Caribbean. In Biogeography of the West Indies: Patterns and Perspectives, edited by C. A. Woods and F. E. Sergile, pp. 481–518. CRC Press, Boca Raton, Florida. 2001b Patterns of Resource Use in the Prehistoric Caribbean. In Proceedings of the Eighteenth International Congress for Caribbean Archaeology, Vol. 2, edited by G. Richard, pp. 78–86. International Association for Caribbean Archaeology, Guadeloupe. 2001c The Sustainability of Resources Used by Native Americans on Four Caribbean Islands. International Journal of Osteoarchaeology 11:112–126. 2009 Zooarchaeology of West Indian Land Mammals. In Terrestrial Mammals of the West Indies, Contributions, edited by R. Borroto-Páez, C. A. Woods, and

322 / References Cited F. E. Sergile, pp. 329–344. Wacahoota Press, Micanopy, Florida, in association with the Florida Museum of Natural History. Witten, Alan J. 2006 Handbook of Geophysics in Archaeology. Equinox, London. Wood, James W., George R. Milner, Henry C. Harpending, and Kenneth M. Weiss 1992 The Osteological Paradox: Problems of Inferring Prehistoric Health from Skeletal Samples. Current Anthropology 33:343–370. Woods, Charles A. 1989 Biogeography of West Indian Rodents. In Biogeography of the West Indies, edited by Charles A. Woods, pp. 741–789. Sand Hill Crane Press, Gainesville, Florida. Yaeger, Jason 2000 The Social Construction of Communities in the Classic Maya Countryside: Strategies of Affiliation in Western Belize. In The Archaeology of Communities, edited by M. A. Canuto and J. Yaeger, pp. 123–142. Routledge, London. Yaeger, Jason, and Marcello A. Canuto 2000 Introducing an Archaeology of Communities. In The Archaeology of Communities, edited by M. A. Canuto and J. Yaeger, pp. 1–15. Routledge, London.

Contributors

Pedro Alvarado Zayas is an archaeologist of the Instituto de Cultura Puertorriqueña in Ponce. He was the field director of the original excavation project at Tibes conducted by the Sociedad Guaynía. His main research interest is Caribbean rock art. Andrew K. Castor currently lives in Salt Lake City and is a staff geologist working for URS Corporation. Melissa J. Castor (née Klinder) lives in Salt Lake City and is currently employed with CH2M Hill as a staff geologist. Edwin F. Crespo-Torres is an assistant professor at the University of Puerto Rico–Río Piedras. He specializes in biological and forensic anthropology. L. Antonio Curet is an associate curator at The Field Museum of Natural History of Chicago. His main interest is the study of social and cultural change, specifically in the pre-Columbian Caribbean and Mesoamerica. Susan D. deFrance is an associate professor at the University of Florida. She conducts research in the Central Andes, the Caribbean, and the southeastern United States. Geoffrey DuChemin is a doctoral student at the University of Florida. His research interest is zooarchaeology of the Caribbean, including the Bahamian islands and Puerto Rico. Richard H. Fluegeman is professor of geological sciences at Ball State University, where he has taught for the past 25 years. Currently he is conducting biostratigraphic research on global change during the Eocene Epoch. Jeffry D. Grigsby is professor of geological sciences and Associate Dean of the College of Sciences and Humanities at Ball State University. He has published in the area of sedimentary petrology/geochemistry.

324 / Contributors Carla S. Hadden works as an archaeological technician for the National Park Service Southeast Archeological Center in Tallahassee, Florida. Her focus is in faunal analysis and environmental archaeology. Michelle J. LeFebvre is a doctoral student at the University of Florida. Her research interest is human use of islands and zooarchaeology in the prehistoric past in the southern Caribbean and the southeastern United States. Lee A. Newsom is associate professor of anthropological archaeology at The Pennsylvania State University. She has published a number of books and articles on the paleoethnobotany of eastern North America and the Caribbean. William J. Pestle is a Ph.D. candidate in the Department of Anthropology at the University of Illinois at Chicago and a Graduate Fellow at The Field Museum of Natural History of Chicago. His current research is focused on the social dimensions of paleodiet in prehistoric Puerto Rico. Scott Rice-Snow is Chair of the Department of Geological Sciences at Ball State University. He is a geomorphologist and hydrologist, with other published Puerto Rican research in morphometry of drainage basins and cave wall alcoves. Lisa M. Stringer is an associate of the Department of Anthropology at The Field Museum of Natural History of Chicago. She has worked on archaeological projects in Peru, Arkansas, Illinois, and Puerto Rico. Joshua M. Torres is a Ph.D. candidate at the University of Florida. His research interests are focused on geographical information systems applications in archaeology, social landscapes, settlement patterns, the archaeology of communities, and the development of complex precolonial social systems of the circumCaribbean. Jeffery B. Walker is the U.S. Forest Service archaeologist of the Caribbean National Forest at El Yunque, Puerto Rico. His research focuses on the analysis of stone artifacts in the Caribbean. Daniel Welch is the manager of the customer-training program for Geophysical Survey Systems, Inc., Salem, New Hampshire. He specializes in the uses of geophysical techniques in archaeology and is a Research Fellow with the Department of Archaeology of Boston University.

Index

Page numbers in italics refer to figures and tables. agriculture, 81–82, 86; home gardens, 81–82, 107, 111; shaman (beheque) as cultivator, 107; special beds, 81–82, 107 agroforestry: forest management, 103, 107, 109, 111–12 allometric regression forumula, 119, 120 Alvarado Zayas, Pedro 14, 16, 19, 24, 178, 179, 193, 265, 278, 280 antemortem: bone fracture, 197, 203; tooth loss (AMTL), 196–97 ancestor worship, 204, 234, 250, 280 apatite, 214, 217, 220, 227 Archaic, 6, 82, 83, 84, 127, 172, 198, 247; interaction with Saladoid groups, 245 artificial cranial deformation, 53; cultural modification, 198, 204–207 auditory exostosis, 203. See also benign bone tumors ball court, 7, 8, 12, 17, 24, 177, 262–63, 183; burials, 30, 37; construction, 21, 32, 45, 189; definition, 20, 264. See also batey ball game, 262, 264, 265 Barinas II, 256 batey, 20, 155, 233, 248; definition, 262; de

la Herradura, 25, 181; de una Hilera, 26; del Cacique, 28–29; del Cemí, 25–26, 174, 181; del Murciélago, 28; Numero 1, 24; Santa Elena, 26, 179. See also ball court; ball game benign bone tumors, 196–97 bipolar reduction techniques, 153, 154, 156, 157, 159, 162, 163, 164, 165, 168, 169, 170, 175; artifacts, 157; core, 156, 157, 159, 162, 165, 167, 175, 176n2; flake, 156, 158, 159, 161, 162, 165, 166, 167, 176n2 birds, 12, 86, 121, 129–30, 132, 136, 143, 150, 222, 223, 224, 225, 228, 229 bone pathologies, 197, 207 boulders, 173, 177 burial, 17, 29–30, 45, 53, 191, 193; clusters, 27, 30, 278; Cuevas, 27, 29; Elenan Ostionoid, 29; Maisabels, 191, 195, 196, 198, 200, 202, 203, 206, 207; Paso del Indio, 191, 195, 196, 198, 199, 200, 201, 202, 204, 206, 207; Punta Candelero, 191, 195, 196, 198, 199, 200, 201, 202, 203, 206, 207; Saladoid, 26, 27, 29–30, 37, 53, 278, 280, 284 button osteoma, 197. See also benign bone tumors

326 / Index C4/CAM plants. See stable isotope cacique, 4, 174. See also chiefdom Caguana, 12, 23, 27, 171, 249, 261, 264, 265, 266, 267, 271, 280, 282 calculus, 194, 196, 202. See also dental calculus canas, 191, 243, 244, 250 Capá style, 8, 53, 153, 239–41, 256, 257. See also Chican Ostionoid; Ostionoid; Taíno Caracoles, 192–93, 285 caries, 194–96, 200, 206 causeway, 21, 23, 25–28, 233, 265. See also calzada; pavement calzada, 20. See also causeway; pavement Cayo Cofresí, 245 Cedrosan Saladoid, 6, 7, 14, 30, 169, 204–205, 206, 209, 210. See also Cuevas style; Hacienda Grande style; Saladoid cemis, 170, 174, 176n3. See also three-pointers ceremonial center, 8, 37, 258, 282–84; definition, 263–64; function, 231, 256, 261, 268–69, 276–77; vacant vs. occupied, 229, 267, 270–72 Cerrillos River Valley, 245, 251, 252, 256–57 Chican Ostionoid, 7–8, 15, 23, 37, 53, 55, 153, 171–72, 175, 198, 209, 235, 255, 258, 281. See also Ostionoid; Taíno chiefdom, 38, 207, 284, 286. See also cacique collagen, 214–15, 220 cobbles, 21, 156, 159, 170, 177 cohoba (Anadenanthera peregrina), 91, 109– 10, 275–76 cojobana (Anadenanthera peregrina) 91–92, 95, 98, 100–101, 105–11. See also cohoba cojobilla. See cohoba; cojobana Collores, 243, 244, 250, 252 complex socialities, 259 cooking, 52, 88, 106, 149; area, 50–52, 55, 72, 88, 89, 149, 166, 170, 174, 274, 280; fuel for, 102, 104; vessels, 271 co-presence, 245 community, 232–33, 238, 243, 245, 249, 250, 254, 284 cost distance model, 236 cribra orbitalia, 203, 211. See also hematological disorder crop and garden plants. See names of plants Cuevas style, 7, 14, 27, 28, 45, 52, 55, 157, 159,

169, 175, 198, 204, 205, 243–45, 278, 280. See also Cedrosan Saladoid; Saladoid demographic profile, 206 dental: caries, 194–96, 200–202, 206; calculus, 194, 196, 202; dental pathologies, 194; pathological conditions, 194–97, 206 dental wear, 194–97, 201 Diego Hernandez, 244 direct freehand percussion, 154, 156, 158, 159, 160, 161–68, 170, 176n2 distribution of: age at death, 194–95, 198–99, 206; sex, 194–95, 198, 206 dog, domestic (Canis familiaris), 122, 132, 135, 138, 144, 150 El Bronce, 83, 172, 192–93, 251, 252, 257, 283 El Cayito, 256 El Colmado Perez, 248, 249, 255 El Fresal, 84 Elenan Ostionoid, 7, 14–15, 23–24, 29, 37–38, 55, 70, 87, 152–53, 156, 168–69, 170, 172– 73, 175, 177, 183, 198, 209, 231, 235, 241, 250, 251–52, 257, 265, 271, 273–74, 278– 82, 285. See also Ostionoid electrical resistivity, 16, 39, 49, 50 62–63, 65, 66, 67, 69, 71, 72, 73, 74, 75, 76, 77, 79; definition, 62 enamel hypoplasia, 200, 203, 211 estimation of height, 194, 198, 200, 206 Esperanza style, 8, 241, 256, 257. See also Chican Ostionoid; Ostionoid; Taíno evening primrose (Oenothera sp.), 84, 90, 91, 94, 96, 102, 105–11, 275 faunal assemblage, 121, 142, 150, 276 fishes: bony, 121, 129, 130–31, 132; cartilaginous, 121, 129, 130–31, 132; reef, 133, 138, 140; riverine, 116, 117, 133, 138, 140, 144, 150 formation process, 30, 267; cultural, 32–33, 277; natural, 30–32 fronto-occipital deformation, 198. See also artificial cranial deformation gender, 17, 152, 154, 163, 165, 169–70, 175 Geographical Information System, 234–35

Index / 327 geomorphological studies, 203 geophysics, 60, 79 González Colon, Juan, 19, 176n1, 176n3, 193 GPR. See ground-penetrating radar ground-penetrating radar (GPR), 62, 71, 72, 74; definition, 72 guanábana (Annona sp.), 90–91, 94, 96, 102, 105, 107, 111 guava (Psidium sp.), 93, 98, 100, 105, 107 guayacán (Guaiacum sp.), 87, 93–95, 98, 100– 101, 103, 105–106 guinea pig (Cavia porcellus), 82–83, 121, 136, 143, 150, 273 Hacienda Grande style, 6, 7, 14, 26, 45, 52, 157, 198, 204, 205, 231, 243, 244, 245, 247, 278. See also Cedrosan Saladoid; Saladoid hematological disorder, 197 Hernández Colón, 244, 251 Huecoid component, 205, 206, 207, 210 hutia, West Indian (Isolobodon portorricencis), 121, 122, 126, 132, 135, 138, 143, 224, 273 inland-coastal interaction, 249, 252 intermediate dimension, 179, 181, 184 interpersonal violence, 203, 207 invertebrates: arks (Arcidae), 125, 133, 136; conchs, 119, 125, 133, 135, 145, 148; mollusks, 116, 117, 121, 136, 144, 148–49, 150, 224, 226, 227, 228, 229; turretsnails (Turritella sp.), 124, 133, 134, 136, 145 IsoSource, 219, 224. See also multisource mixture modeling isotope. See stable isotopes Jacana, 154, 171, 172, 281. See also PO-29 jagua (Genipa americana), 93, 105, 107–108 jobo (Spondias sp.), 92–93, 105, 107 La Hueca, Site, 205; Style, 6, 7, 198, 247 La Mineral, 239, 248 Lago Gely, 241, 251 Land crab, 124, 133, 224 Las Flores, 241, 244, 249, 252, 257, 283 lingual surface attrition of maxillary anterior teeth (LSAMAT), 196, 202, 207 lithology, 2, 23, 58, 177–85

LSAMAT. See Lingual surface attrition of maxillary anterior teeth Luna Calderón, Fernando, 193, 204, 207 Macroblock, 39, 41, 44, 46–49, 51, 88–91, 94, 97, 101, 104, 118–19, 121, 127, 130–31, 161–66, 173–74 maga (Thespesia grandiflora), 93, 94, 98, 100 magnetic gradiometry, 16, 41, 50, 71, 73, 76 Maisabel, 191, 198, 203, 204, 207, 209; burials, 191, 195, 196, 198, 200, 202, 203, 206, 207; cranial deformation, 204, 207; stable isotope analysis, 201 maize, 11, 81, 84, 111, 113n1, 200–201, 206, 213, 227 mammals. See guinea pigs; hutia; marine mammals manioc (Manihot esculenta), 11, 84, 113n1, 196, 200, 201, 206, 210; grater, 153–54, 163, 166, 167, 169 marine mammals, 132, 136, 144; manatee (Trichechus sp.), 122, 135, 138; seal (Monachus sp.), 222, 223, 224, 228 Maruca, 84, 193, 245, 215, 216, 220, 229, 245 microfossils, 187–88 Mona Passage, 250 Monserrate style, 7, 14–15, 28, 45, 55, 90, 119, 127, 156, 159, 273, 279, 282. See also Elenan Ostionoid; Ostionoid; Pre-Taíno Morel site, Guadaloupe, 205 multisource mixture modeling, 211, 219, 224, 225 narcotic plants, 81, 82, 84, 93, 102, 104, 105, 106–10, 275 níspero (Manilkara zapota), 90–91, 94, 107 nutritional stress conditions, 203, 207, 211 OP19, 39, 41, 47, 88, 94, 97, 104, 118, 130, 166–67, 174, 274; OP19A, 69; OP19B, 51–52, 70, 97, 99, 101, 162, 167; OP19C, 51, 70, 88, 94, 97, 99, 100, 101, 119, 127, 162, 167, 168, 176n2; OP19D, 88, 97, 101, 104, 119, 127, 131, 162, 167; OP19E, 55, 56, 97, 100, 101, 118, 119, 120, 127, 133, 134–36, 145, 148–49, 162, 167; OP19F, 101, 113n4, 162; OP19G, 101, 113n4, 162,

328 / Index osteoarthritis, 196, 197, 207 osteoarticular lesion, 197. See also osteoartrhritis Ostionoid 7, 43, 80, 83, 84, 87, 89–90, 100, 102, 108, 111, 113n1, 115, 118, 119, 121 122– 26, 127–28, 129–31, 132–35, 136–37, 143, 145, 148–51, 164, 169, 198, 204, 233, 235, 244, 245, 250–51, 279–80. See also Chican Ostionoid; Elenan Ostionoid; Ostionan Ostionoid Ostionan Ostionoid, 7, 15, 87, 177, 235, 250, 252, 257, 280, 281, 285. See also Ostionoid; Pre-Taíno Ostiones styles, 7, 53, 90, 158–59, 161, 241, 245, 251–52; Modified, 7, 53, 90, 158, 161; Pure, 7, 90, 159. See also, Ostionan Ostionoid; Ostionoid; Pre-Taíno paleodemography models, 199; study, 198 papaya (Carica papaya), 84 parrot (Amazonas sp.), 122, 132, 138, 140, 143 Parry fracture, 203 Paso del Indio, 113n1, 144, 198, 215; burials, 191, 195, 196, 198, 199, 200, 201, 202, 204, 206, 207; cranial deformation, 203, 207; stable isotope analysis, 113n1, 201, 216, 220 pathological conditions, 194–98, 202, 203, 206, 207 pavement, 17, 155, 161, 171, 173, 175, 177, 178–82, 183–85, 186, 188. See also calzada perimortem trauma, 204 periodontitis, 196–97 periostitis, 196–97, 202–203, 207 petroglyph, 17, 22, 27, 153, 154, 156, 170, 171, 172, 174, 178, 184, 189, 236, 249, 265, 266 pinnipeds, 222, 223, 224, 225, 228, 283. See also marine mammals plaza, 7, 8, 17, 20, 37, 42, 69, 87, 150, 152, 154, 155, 156, 168, 169, 171, 172, 175, 176n1, 177–79, 181–85, 189, 233, 234, 236, 239–41, 247, 249, 262–67, 271, 279, 284, 286; construction, 20–21, 31, 32, 45, 170–73, 175, 178–79, 183–85; definition, 20, 262–64; Plaza de la Estrella, 21, 24, 27–28, 45, 46, 66, 74, 78, 174, 175, 179,

181, 182, 183, 193; Plaza Principal, 21, 24, 26–27, 28, 29, 46, 53, 63, 66, 68, 72, 158, 161, 166, 168, 174, 185, 193, 253, 274, 278, 280, 281 PO-21, 84, 241, 251, 253 PO-23, 241, 245, 251, 253 PO-27, 241, 253, 256, 257 PO-29, 172, 240, 248, 249, 251, 253, 255, 256. See also Jacana PO-38, 84, 239, 245, 251, 252, 253 PO-39, 84, 239, 252, 253, 256, 283 PO-42, 248, 249, 251, 252, 253, 255. See also La Mineral PO-43, 248, 249, 251, 255 pollen analysis, 32, 81, 89 porotic hyperostosis, 197, 200, 203. See also hematological disorder Portugués River, 8, 12, 19, 23, 26, 31, 66, 77, 85, 117, 133, 144, 149, 155, 156, 170, 172, 177–81, 183, 184, 189, 226, 248, 249, 255, 256, 258, 259, 280, 281, 285 Pre-Taíno, 155, 177, 187, 188, 266. See also Elenan Ostionoid; Monserrate style; Ostionan Ostionoid; Ostiones style; Santa Elena style Punta Candelero, 191, 198, 215, 220; burials, 191, 195, 196, 198, 199, 200, 201, 202, 203, 206, 207; cranial deformation, 204, 207; stable isotope analysis, 201, 216, 220 radiocarbon dates, 8, 14, 15, 36, 39, 53, 55–58, 70, 118, 153, 156, 158, 164, 169, 235, 236, 241, 243, 245, 246, 248, 252, 253, 254, 255, 257, 266, 277, 278, 279, 280, 281 reptiles, 12, 121 129, 130; iguanas, 224; lizards, 121, 122, 132, 136, 138, 150. See also turtles rodents, 25, 82, 122, 126, 132, 138, 143, 150, 273. See also guinea pigs; hutias root crops, 82, 84, 111, 113n1, 117, 196 Rouse, Irving, 204, 235, 268 Saladoid, 7, 14, 29, 37, 38, 41, 43, 52, 53, 54, 55, 80, 84, 87, 89, 90, 100, 102, 106, 108, 111, 112, 115, 118, 119, 121, 122–26, 126, 127–30, 132, 133, 134, 135, 136, 137, 143, 144, 145, 148, 149, 150–51, 153, 154, 157, 161, 164, 167, 168, 169, 172, 173, 177,

Index / 329 198, 204, 205, 207, 232, 233, 235, 243, 244, 245, 247, 248, 250, 251, 254, 258, 265, 273, 274, 275, 278, 279, 280, 281; burials, 26, 27, 29, 30, 37, 53, 278, 280, 284; veneer, 245–47, 254, 258. See also Cedrosan Saladoid sandstone, 10, 155, 156, 158, 165, 166, 167, 168, 171, 172, 180, 181, 182, 184–85, 187, 188, 189 Santa Elena style, 7, 14, 15, 24, 26, 27, 28, 45, 49, 53, 89, 90, 119, 127, 137, 156, 158, 159, 161, 164, 175, 204, 251, 257, 282. See also Elenan Ostionoid; Ostionoid; Pre-Taíno settlement types, 235–36 sex ratios, 200, 206 Shannon-Weaver index, 120, 121, 145 Sheldon equitability index, 120, 121, 145 shovel-shape incisors, 197, 202, 207 Sociedad Arqueológica del Sur-Oeste de Puerto Rico, 19, 33 Sociedad Guaynía de Arqueología e Historia de Ponce, 16, 19 20, 29, 33, 34, 37, 69, 178, 265, 271, 278, 280 Sonadora, 239, 256, 266 stable isotope, 17, 201, 211, 216–18, 221, 229; C3 plants, 201, 213, 222, 227; C4/CAM plants, 113n1, 201, 213, 214, 215, 219, 221, 222, 227, 228, 229; dietary reconstruction, 212–15 stone collar, 153, 154, 168–70, 175, 176n3

stone rows, 20, 24, 25, 26, 27, 28, 32, 34, 233, 263; construction, 20, 21–24, 32 subtropical dry forest ecosystem, 10, 85–87, 91, 102, 103, 105; deciduous forest, 85–87, 94, 103; evergreen forest, 85, 103; mangrove forest, 8; riparian forest, 86; salt flat and beach thicket, 85; scrub forest, 85; semievergreen forest, 85 subtropical moist forest, 11 subtropical wet forest, 11 Tachuelo (Pictetia aculeata), 87, 91, 92, 94, 98, 100, 103, 105, 113 Taíno, 7, 37, 82, 107, 110, 262. See also Chican Ostionoid taphonomy, 88, 89, 194, 204, 220 Tecla, 239, 243, 244, 245, 250, 252, 253 three-pointers, 153, 154, 168, 169, 175, 272. See also cemis tobacco (Nicotiana sp.), 109, 110 tooth loss, antemortem (AMTL), 196–97 traumas, 196–97, 203 204, 207 turtles: marine (Cheloniidae), 122, 132, 135, 139, 144, 222, 225; riverine, 132 vacant ceremonial centers, 229, 265, 266, 267, 269, 270–76 vertebrates, 16, 116, 121, 127–31, 150 zooarchaeology, 86, 118, 119, 210