ANCIENT MAYA DIET AT CALEDONIA, CAYO DISTRICT, : THE ISOTOPIC EVIDENCE

A Thesis Submitted to the Committee of Graduate Studies in Partial Fulfillment of the Requirements for the Degree of Master of Arts in the Faculty of Arts and Science.

TRENT UNIVERSITY Peterborough, Ontario, Canada

© Copyright by Asta J. Rand 2011

Anthropology M.A. Graduate Program

January 2012 Library and Archives Bibliotheque et 1*1 Canada Archives Canada Published Heritage Direction du Branch Patrimoine de I'edition

395 Wellington Street 395, rue Wellington Ottawa ON K1A 0N4 Ottawa ON K1A 0N4 Canada Canada

Your file Votre reference ISBN: 978-0-494-82186-2 Our file Notre reference ISBN: 978-0-494-82186-2

NOTICE: AVIS:

The author has granted a non­ L'auteur a accorde une licence non exclusive exclusive license allowing Library and permettant a la Bibliotheque et Archives Archives Canada to reproduce, Canada de reproduire, publier, archiver, publish, archive, preserve, conserve, sauvegarder, conserver, transmettre au public communicate to the public by par telecommunication ou par I'lnternet, preter, telecommunication or on the Internet, distribuer et vendre des theses partout dans le loan, distribute and sell theses monde, a des fins commerciales ou autres, sur worldwide, for commercial or non­ support microforme, papier, electronique et/ou commercial purposes, in microform, autres formats. paper, electronic and/or any other formats.

The author retains copyright L'auteur conserve la propriete du droit d'auteur ownership and moral rights in this et des droits moraux qui protege cette these. Ni thesis. Neither the thesis nor la these ni des extraits substantiels de celle-ci substantial extracts from it may be ne doivent etre imprimes ou autrement printed or otherwise reproduced reproduits sans son autorisation. without the author's permission.

In compliance with the Canadian Conformement a la loi canadienne sur la Privacy Act some supporting forms protection de la vie privee, quelques may have been removed from this formulaires secondaires ont ete enleves de thesis. cette these.

While these forms may be included Bien que ces formulaires aient inclus dans in the document page count, their la pagination, il n'y aura aucun contenu removal does not represent any loss manquant. of content from the thesis.

1*1 Canada ABSTRACT

Ancient Maya Diet at Caledonia, Cayo District, Belize: The Isotopic Evidence

Asta J. Rand

Ancient Maya diet at Caledonia, Cayo District, Belize (Late Preclassic to

Terminal Classic periods), was investigated using stable carbon and nitrogen isotope analysis of human bone collagen, bioapatite and tooth enamel bioapatite. These data were compared to skeletal and dental evidence of diet and nutritional stress, and dietary differences based on age, sex, social status and time period were considered. The diet of the Caledonia Maya was based on maize, supplemented with maize-fed terrestrial animal protein and freshwater mollusc protein. Due to small sample sizes, no relationships between diet and pathology, age, sex, social status or time period were discerned. The

Caledonia Maya did consume diets similar to those at neighbouring sites, as well as to sites in the Peten of . This study demonstrates that although small sample sizes can preclude the detection of intra-site trends at minor centres such as Caledonia, inter- site comparisons can provide information on regional dietary practices.

Keywords: Ancient Maya, Caledonia, Belize, paleodiet, stable isotopes, carbon, nitrogen, collagen, bioapatite, human bone, human teeth ACKNOWLEDGEMENTS

So many people have provided me with guidance, encouragement and support

over the last two years, and I would like to take the opportunity to express my gratitude.

First, of course, I would like to thank my supervisor, Dr. Anne Keenleyside, who has

been there from my undergraduate years at Trent through to the completion of this thesis.

Without your tireless support, encouragement and advice, I do not think I would have

applied to graduate school, much less completed a Master's thesis.

Second, I would like to thank the members of my committee, Drs. Jocelyn

Williams and Paul Healy; you both have done so much to make this research possible. Dr.

Healy not only provided me with the skeletal sample analyzed in this thesis, and photos of

Caledonia during excavation, but also inspired my interest in the ancient Maya. You were

always there to answer a question when I asked and I am forever grateful for your help

and support. Dr. Williams, I thank you for lending me your M.A. thesis, which was a

great asset to this project, and without your assistance I would still be painstakingly

interpreting my data. I greatly appreciate your patience, encouragement and all of the help you have provided over the years. I would also like to thank Dr. Lori Wright for her role

as external examiner during my defence, and the helpful comments she provided.

I would like to thank the Trent Research Ethics Board for granting approval for the research on human skeletal remains from Caledonia, Cayo District, Belize. I would

also like to thank Dr. Jaime Awe, who granted permission from the government of Belize

for this skeletal sample to be analyzed, and whose M.A. thesis was a priceless asset to my own research. Dr. Henry Schwarcz also provided valuable insights into isotopic and

FTIR analysis through Dr. Keenleyside, without which I would not have managed. I will

iii be forever grateful to Martin Knyf, who ran my samples so quickly at McMaster

University and answered any and all questions I had about the process. I am also grateful to the Social Sciences and Humanities Research Council (SSHRC) for the Master's scholarship I received from the Joseph-Armand Bombardier Canada Graduate

Scholarships program (Award no: 766-2010-0718) in September 2010. Without SSHRC funding, this research would never have been possible.

Everyone at Trent was exceptionally supportive and kind. Kristine Williams was always there when I needed to chat and has provided so much support over the last few years. Also, the storage of the Caledonia collection would not have been possible without the help of Kate Dougherty. I would also like to thank Barry Best in the Chemistry

Department at Trent for his assistance with the FTIR analysis of carbonate in the bone and tooth bioapatite samples in this study. Dr. James Connolly also provided invaluable help with my understanding of the statistical analyses used in this study.

I owe a great deal of gratitude to all of my fellow anthropology graduate students at Trent University. You were a wonderful group of people, and the last two years would not have been the same without all of you. There are several people I am especially indebted to: Megan Bower, Madeline Holder, Kimberly Jankuta, and Cara Tremain, among others, who provided helpful comments and suggestions on several of my assignments and presentations over the last two years. Also, without Kimberly's help I would have never completed the statistical analysis in this study. I would also like to thank Max Lamoureux St.-Hilaire for lending me several books on the ancient Maya and helping me to better understand the archaeological context of the Caledonia sample.

Finally, I would never have survived the last two years without Megan. Your friendship, enthusiasm and long chats got me through many bumps along the road, although I am still

iv not sure about those aliquots.

I would not have remained sane these last few years without the friendship of so

many people. Murphy and Ron have been there since I began graduate school, and my

weekends would not have been the same without them. Rhiannon provided invaluable

advice when I first began writing that I'm sure has saved me much grief. I would also like to thank Bill, Danno, Steve and all who no longer live in Peterborough, but who keep

coming back to visit and put a smile on my face.

Finally, I am exceedingly grateful to my whole family, especially my parents,

Leah and Paul, and my siblings, Aubrey and Casey. Your tireless love and

encouragement, as well as your enthusiasm regarding my research have meant so much to me. And of course I thank Drew; you have been my constant throughout this whole process. I do not know how I would have made it these last six years without your love and support, let alone to the completion of this thesis. I dedicate this thesis to my family and to Drew.

vi TABLE OF CONTENTS

Abstract ii

Acknowledgements iii

Dedication vi

Table of Contents vii

List of Figures xi

List of Tables xiii

Chapter One: Introduction 1 1.1. The Ancient Maya 2 1.2. Thesis Overview .4

Chapter Two: Archaeological Context of the Sample 7 2.1. Site Location and Geography 7 2.2. Archaeological Investigations and Burial Descriptions 9 2.2.1. Plaza A - Burials # land #2 11 2.2.1.1. Burial #1 13 2.2.1.2. Burial #2 13 2.2.2. Plaza C-Burials #3, #4 and #5 14 2.2.2.1. Burial #3 14 2.2.2.2. Burial #4 15 2.2.2.3. Burial #5 16 2.2.3. Additional Artefactual and Faunal Remains 16

Chapter Three: Theoretical Principles of Stable Isotope Analysis 17 3.1. The Principles of Stable Isotopes 17 3.2. Stable Carbon Isotopes 19 3.2.1. Stable Carbon Isotopes and Dietary Reconstruction 19 3.2.2. Carbon in Terrestrial Ecosystems 19 3.2.3. Carbon in Marine Ecosystems 21 3.2.4. Carbon in Freshwater Ecosystems 22 1 "^ 3.2.5. Complicating Factors that Influence 8 C Values 22 3.3. Stable Nitrogen Isotopes 24 3.3.1. Stable Nitrogen Isotopes and Dietary Reconstruction 24 3.3.2. Nitrogen in Terrestrial Ecosystems 25 3.3.3. Nitrogen in Aquatic Ecosystems 25 3.3.4. Complicating Factors that Influence 815N Values 26 3.4. Bone Biology and Diet 28 3.4.1. Bone Remodelling and Turnover Rates 29 3.4.2. Dietary Fractionation of Carbon and Nitrogen 31

vn Table of Contents Continued

3.4.3. Distribution of Carbon from Macronutrients to Collagen and Bioapatite 32 3.5. Diagenesis 35 3.5.1. Assessing the Integrity of Preservation of Collagen 3 7 3.5.2. Assessing the Integrity of Preservation of Bioapatite 39 3.6. Chapter Three Summary 41

Chapter Four: Ancient Maya Food Resources 43 4.1. Indirect Evidence of Ancient Maya Food Consumption 43 4.1.1. Floral Resources 43 4.1.2. Faunal Resources 47 4.1.3. Skeletal and Dental Pathology 50 4.2. Food Resources at Caledonia 54 4.2.1. Modern Subsistence Resources at Caledonia 54 4.2.2. Archaeological Evidence of Dietary Resources at Caledonia 55 4.3. Stable Carbon and Nitrogen Isotope Studies in the Maya Region 57 4.3.1. Isotopic Food Webs in Mesoamerica 57 4.3.2. Stable Isotope Analysis of Ancient Maya Diet Review 62 4.3.2.1. Variability by Age 64 4.3.2.2. Variability by Sex 65 4.3.2.3. Variability by Social Status 66 4.3.2.4. Temporal Variability 69 4.3.2.5. Regional Variability in Diet 70 4.4. Chapter Four Summary 71

Chapter Five: Analytical Techniques 73 5.1. Methods Employed in Skeletal Analysis 73 5.1.1. Age Estimation Techniques 74 5.1.1.1. Age Estimation Techniques for Subadult Remains 74 5.1.1.2. Age Estimation Techniques for Adult Remains 75 5.1.2. Sex Determination Techniques 77 5.1.3. Pathology: Nutrition and Disease 78 5.2. Preparation of Bone Samples for Stable Isotope Analysis 79 5.2.1. Collagen 79 5.2.2. Bone Bioapatite 84 5.2.3. Tooth Enamel Bioapatite 86 5.2.4. FTIR Analysis 87 5.2.5. Sample Analysis 88 5.2.6. Statistical Analysis 88

Chapter Six: Results 90 6.1. Skeletal Analysis: Age, Sex, and Pathology 90 6.1.1. Burial #1 90 6.1.2. Burial #3 91 6.1.2.1. Individual A 91

vm Table of Contents Continued

6.1.2.2. Individual B 92 6.1.2.3. Individual C 92 6.1.2.4. Individual D 93 6.1.3. Burial #4 94 6.1.3.1. Individual A 94 6.1.3.2. Individual B 94 6.1.3.3. Individual C 95 6.1.3.4. Individuals D and E 95 6.1.3.5.. Fibulae 95 6.1.4. Burial #5 Individual A 96 6.1.5. Unknown Burial 96 6.2. Sample Preservation 96 6.2.1. Preservation of Collagen 96 6.2.2. Preservation of Bioapatite 98 6.3. Stable Nitroj*e n and Carbon Isotope Data 99 6.3.1. Summary of Isotope Data 99 6.3.2. Variability by Age 102 6.3.3. Variability by Sex 107 6.3.4. Temporal Variability 110 6.3.5. Variability by Burial Type 113 6.4. Chapter Six Summary 116

Chapter Seven: Discussion and Interpretations 117 7.1. Interpretation of Isotopic Analysis 117 7.1.1. General Diet at Caledonia, Belize 117 7.1.2. Variability by Age 127 7.1.3. Variability by Sex 128 7.1.4. Temporal Variability 130 7.1.5. Variability by Socioeconomic Status 132 7.1.6. Regional Comparison of Ancient Maya Diet 137 7.2. Health, Pathology and Isotope Analysis 146 7.2.1. Porotic Hyperostosis 147 7.2.2. Dental Pathology 151 7.2.2.1. Linear Enamel Hypoplasia 152 7.2.2.2. Dental Calculus 153 7.2.2.3. Dental Caries 154 7.2.2.4. Dental Wear 156 7.2.3. Pathological Summary 157 7.3. Methodological and Theoretical Concerns 157 7.3.1. Sample Size 157 7.3.2. Representativeness of the Sample 160 7.3.2.1. Uncontrollable Factors 160 7.3.2.2. Controllable Factors 164 7.4. Chapter Seven Summary 165

IX Table of Contents Continued

Chapter Eight: Conclusions 167 8.1. Summary and Significance of Research 167 8.2. Limitations of this study 169 8.3. Future Research 171

References Cited 175

Appendices Appendix A: Non-Human Skeletal Remains and Artefacts 208 Appendix B: Plant and Animal Remains Recovered from Ancient Maya Sites 211 Appendix C: Flora and Fauna Stable.Carbon and Nitrogen Isotope Data Used to Create Food Web 221 Appendix D: Mean Stable Isotope Data for Ancient Maya Skeletons 231 Appendix E: Porotic Hyperostosis and Dental Disease in the Caledonia Skeletal Sample 236 Appendix F: Results of Age, Sex and Stature Estimates 241 Appendix G: Summary of Isotope Results 244 Appendix H: Results of Statistical Analyses 246

x LIST OF FIGURES

Figure 1.1 Map of the Maya region, including regional divisions and sites investigated using stable isotope analysis 3 Figure 3.1 The isotopic composition of an element 18 Figure 3.2 Equations used to calculate collagen yield 37 Figure 3.3 Typical FTIR graph and method for calculating crystallinity index (CI) (after Shemesh 1990:2434; Weiner and Bar-Yosef 1990:191) 41 Figure 4.1 Deer depicted on Vessel 8 recovered from Burial #1 in Str. A-1 from Caledonia, dating to the Early to Late Classic transition (A.D. 450 to 650) (after Awe 1985:138-140; Healy et al 1998:267) 56 Figure 4.2 Food webs for two different areas occupied by the ancient Maya. Sites from inland Belize (Coyston et al. 1999:225) are illustrated on the left, and sites from the Pasion Valley region of Guatemala (Wright 1997a: 187) are illustrated on the right. 59 Figure 4.3 Composite food web of inland and coastal resources that would have been available to the ancient Maya at Caledonia (data from Keegan and DeNiro 1988; Norr 1991 White and Schwarcz 1989; White et al 1993; Williams 2000; Williams et al. 2009; Wright 2006) 60 13 15 Figure 6.1 8 Ccoi and 8 N values for 20 sampled individuals 100 1 3 Figure 6.2 8 Cbi0 results for 18 sampled individuals 100 13 Figure 6.3 A Ct,i0-coi values for 18 sampled individuals 101 Figure 6.4 8 Cen results for 6 sampled individuals (outlier sample from C2-3-A is circled) 101 13 15 Figure 6.5 8 Ccoi and 8 N results by age group 104 13 Figure 6.6 8 Cbi0 results by age group 104 13 Figure 6.7 A Cbi0 results by age group 106 1 3 Figure 6.8 8 Cen results by age group 107 13 13 Figure 6.9 Difference between 8 Cb,0 and S Cen values 107 13 15 Figure 6.10 8 Ccoi and 8 N results by sex 108 13 Figure 6.11 8 Cbi0 results by sex 109 1"} Figure 6.12 A Cbi0-coi results by sex 109 13 Figure 6.13 8 Cenbysex 109 13 15 Figure 6.14 8 Ccoi and 8 N results by time period 110 13 Figure 6.15 8 Cbi0 results by time period 111 1 3 Figure 6.16 A Cb,0-Coi results by time period 112 1 "\ Figure 6.17 8 Cen results by time period 112 13 15 Figure 6.18 8 Ccoi and 8 N results by burial type 113 1 "3 Figure 6.19 8 Cbi0 results by burial type 114 Figure 6.20 A Cbio-coi results by burial type 115 13 Figure 6.21- 8 Cen results by burial type 115 Figure 7.1 Foodweb for Caledonia (see Section 4.3.1 for details), illustrating uncorrected human collagen values (X) 118 Figure 7.2 Caledonia data plotted according to Krueger and Sullivan (1984) 124 Figure 7.3 Caledonia data plotted according to Kellner and Schoeninger (2007). 13 Individuals that exhibit A Cbi0-coi values below 3%o are circled. 124

XI List of Figures Continued

Figure 7.4 Average 8 Ccoi and 8 Cbio values for ancient Maya sites plotted according to Kellner and Schoeninger (2007). Sites where individuals are known to have consumed fish are circled 126 13 15 Figure 7.5 Mean 8 Ccoi and S N values for ancient Maya sites located in Inland Belize (X), coastal Belize (•), Lowland Guatemala (+), highland Guatemala (-), (A), and (x). Caledonia is circled 139 11 1 ^ Figure 7.6 Mean 8 Ccoi and 8 Cbi0 values for skeletons from ancient Maya sites located in inland Belize (x), coastal Belize (•), Lowland Guatemala (+), Honduras (A), and Mexico (x). Caledonia is circled 141

xn LIST OF TABLES

Table 2.1 Time periods used at Caledonia 9 Table 2.2 Summary of burials from Caledonia 12 Table 5.1 Tissues sampled for stable isotope analysis 80 Table 6.1 Indicators of sample preservation 97 Table 6.2 Individual results of isotopic analysis 103 Table 6.3 Isotopic summary statistics by sex, age, time period and burial type 105 Table 7.1 Mortuary data for Caledonia burials 134 Table 7.2 Data for individuals with and without porotic hyperostosis 150 Table 7.3 Caries frequencies by tooth count from selected ancient Maya sites 155 Table 7.4 Number of individuals sampled for isotopic analysis from ancient Maya sites 158

Xlll 1

CHAPTER ONE

INTRODUCTION

Food is an important component of all cultures, not just for subsistence purposes, but also in terms of establishing and maintaining social identity (Gumerman 1997; van der Veen 2003). Furthermore, understanding the diet of a past population can provide information about their food procurement techniques, economic relationships, as well as their health. Traditionally, indirect lines of evidence provided by archaeological, ethnographic, ethnohistoric, linguistic, palaeobotanical, palaeopathological, and zooarchaeological studies have been used to indicate the types of foods that were available to the population of interest (i.e., the menu). Although these methods of inferring diet are invaluable for establishing the context in which food resources were selected, they cannot directly indicate the diet of a single individual at a site. In conjunction with these studies, stable carbon and nitrogen isotope analysis of human dental and skeletal remains have the potential to assess directly the types of foods an individual consumed during their life (i.e., the meal). When the results of such analysis are situated within the greater dietary context established using indirect lines of evidence, dietary variability in terms of age, sex, social status, time period and region can be explored.

This thesis employs stable carbon and nitrogen isotope analysis of human skeletal and dental tissue to investigate the diet of the ancient Maya interred at Caledonia, a minor centre located in the Cayo District of Belize, and dating from the Late Preclassic to the

Terminal Classic periods. Archaeological investigations at this site were undertaken in

1980 and again in 1984 (Awe 1985), and the skeletal remains uncovered during these 2

excavations have been previously analyzed (Healy et al. 1998; Helmuth 1985). There is,

however, a dearth of dietary evidence at this site; no palaeobotanical studies have been

undertaken, and very few faunal remains and artefacts related to diet were recovered

during the excavations. The use of stable carbon and nitrogen isotope analysis to

reconstruct diet has been well established in the Maya region (see Chapter Four). Thus,

stable isotope analysis has the potential to provide inferences regarding social, temporal

and regional aspects at Caledonia.

1.1 The Ancient Maya

The ancient Maya had a highly complex and stratified society, and inhabited the

southeastern portion of Mesoamerica from the Preclassic to Historic Period (see Chapter

2 for details on time periods). This area is divided into three regions: the Highlands, which include the southern portions of modern Guatemala and western ; the

Southern Lowlands, including western Honduras, the Peten region of Guatemala, and

Belize; and the Northern Lowlands which includes of the Mexican states of Campeche,

Quintana Roo and Yucatan (Figure 1.1).

Detailed descriptions of the ancient Maya can be found elsewhere (McKillop

2006; Sharer and Traxler 2006), and are briefly summarized below. The Maya were not unified under one empire as was the case with the Aztecs of central Mexico or the Inca of the South American Andes, and instead consisted of many city-states with complex intra- and inter-site political, economic, and cultural relationships. These polities shared a

similar culture, including but not limited to ceramics, artwork, architecture, beliefs, and

staple crops. It was once believed that the ancient Maya subsisted primarily on maize

(corn) with little intake of other food resources. Although the Mesoamerican triad of Ancient Maya Sites I. ChuiK-hucniil 2<*¥axunr 3. Cuelln 4. K'axol) 5. Sao Pedro 6. Mateo Gonzalez 7. Laimtnai kChanHiix 9. Alton Ha W. Mono Cay II. 12. Barton Ramie 13. Pacbiton 14. Cabal Pech 15. Caledonia ,2 U. Csracol 17. Casactaja (7 1$. Hobnul Northern Lowlands 19. Piedras Hegras 20. Itean 21. 22. 23. 24. Agnateca 25. L'opaa MEXICO 26. Iximclij' 27. KaminaTjuvu ^*>

3 4 rJ;r ' »MJ A J .u 6° / Southern Lowlands 1 18> n f f irs. uli \ ^-v •15 f \ 20 *16 J \J" * .22 j BELIZE / ^'23 '24 1 ^

^\ GUATEMALA

Highlands •j5 HONDURAS

.26 *27 N T EI, SALVADOR i Kilometers 0 25 50 100 150 200

I Miles 0 25 50 100 150 200 The Electronic Atlas of Ancient Maya Sites (c) copyright 2008 Clifford T. Brovwi & Wlater f?. T Witschey

Figure 1.1: Map of the area inhabited by the ancient Maya, including regional divisions and sites investigated using stable isotopic analysis. 4 maize, beans and squash was certainly important, and these crops were consumed by the ancient Maya, it has since been established that the Maya exploited a variety of food resources from a range of ecological niches, including aquatic, marine, pine forest and tropical species of both plants and animals (see Chapter Four).

1.2 Thesis Overview

This thesis focuses on the skeletal remains recovered from Caledonia during previous archaeological investigations. The skeletal material from Burial #1 at Caledonia was recovered during the 1980 field season, whereas the remaining skeletons from

Burials #3, #4 and #5 were recovered in 1984 (Awe 1985). The skeletal material was not evaluated in the field, but instead was transported to Trent University where it was analyzed by Dr. Herman Helmuth (Helmuth 1985), and where it is currently curated.

Although the skeletal remains of a subadult from Burial #2 were recorded in the field

(Awe 1985), they were not noted by Helmuth (1985) and are thus not considered in this thesis. Originally, Helmuth (1985; Healy et al. 1998) identified 18 individuals based on reoccurring diagnostic skeletal elements. I reassessed the osteological material from

Caledonia as part of this thesis research in order to ensure that age and sex estimates were consistent for the statistical analyses of isotopic differences between sampled individuals.

Differences between the initial osteological investigations and my own are attributed to the long period in which this collection has been in deep storage, and the fact that this collection has been moved multiple times (see Chapter Six).

Following the osteological analysis, 21 human bone collagen samples were prepared for stable carbon and nitrogen isotope analysis, and 18 bone and 6 tooth enamel bioapatite samples were prepared for stable carbon isotope analysis. Carbon in bone collagen provides information on the protein component of the diet, whereas carbon in

bioapatite reflects the carbon consumed from the whole diet. Nitrogen in bone collagen

can also provide information on protein consumption. It is thus important to sample both

collagen and bioapatite, as well as both carbon and nitrogen in order to provide a better

understanding of the types of resources that were consumed (see Chapter 3). The goals of

this research were as follows:

1) Confirm the biological profiles of the individuals buried at Caledonia established

by Helmuth (1985; Healy et al. 1998).

2) Analyze stable carbon and nitrogen isotopes from a sample of individuals interred

at the site in order to investigate the types of resources that they consumed.

3) Determine whether dietary patterns related to several factors, including age, sex,

social status and time period were present at the site.

4) Investigate the relationship in the past between diet, nutrition, health and disease

at Caledonia.

5) Compare the diet of the Caledonia Maya to that of individuals at other ancient

Maya sites, and consider possible explanations for regional variability in diet

among the Maya.

This thesis is divided into a series of chapters in order to provide context for the

interpretations of the isotopic results. Chapter Two discusses information on the site of

Caledonia, including the environmental setting, the site chronology, important

archaeological features, and other relevant information. A theoretical background on the principles of stable isotope analysis is presented in Chapter Three. This includes a

discussion of what stable isotopes are, how they can be used to discern different dietary components, and how postmortem changes to the chemical composition of bone and tooth

(i.e., diagenesis) can be evaluated and controlled.

Chapter Four explores the floral and faunal resources that would have been consumed by the Maya and available in the area surrounding Caledonia both in the past and present. The chapter concludes with a review of stable isotope studies that have been conducted on human skeletal remains from ancient Maya sites. Chapter Five outlines the analytical techniques employed in this research. First, the methods used to estimate age and sex, and identify pathology are discussed. This is followed by an explanation of the collagen and bioapatite extraction procedures, methods used to evaluate diagenesis, and statistical analyses used in this research.

The results are presented in Chapter Six, beginning with the results of the age and sex estimations, followed by identified pathological conditions, and the stable isotope results. These are discussed in greater detail in Chapter Seven. The general diet of the

Caledonia Maya as inferred from isotopic analysis is explained, including intra-site dietary variability by age, sex, social status and time period. The isotopic results from the individuals interred at Caledonia are then compared with those from other ancient Maya sites, and potential explanations for the regional variability in diet are presented. The relationship between health, pathology and isotope analysis at Caledonia is then explored.

The chapter concludes with a review of theoretical and methodological concerns. Finally,

Chapter Eight provides a summary of the results and a discussion of the significance of this research, a reiteration of the limitations of this study, and concludes with suggestions for future research. 7

CHAPTER TWO ARCHAEOLOGICAL CONTEXT OF THE SAMPLE

2.1 Site Location and Geography

Caledonia is a small Maya site located on the Vaca Plateau within the Chiquibil

Forest Reserve in the southern part of the Cayo District in Belize. The site is situated on a low hill on the west bank of the Macal River roughly 550 m above sea level.

Ecologically, Caledonia is located at the intersection of the limestone-based Vaca Plateau and the granitic, pine-forested Pine Ridge region (Awe 1985:15).

The climatic pattern for Belize in general can be divided into a wet season from

May to January, followed by a dry season. The mean annual rainfall in this region is

560.1 mm (61.43"), with a mean number of rainy days equalling 153.9 days (Johnson and

Chaffey 1973:9-11). The mean annual is a maximum of 29°C (84°F), with a drop to as low as 6°C (42°F) in January and February. The humidity varies from 70% in

May to 90% in January, with a mean annual humidity of roughly 82% (Johnson and

Chaffey 1973:11).

The two major waterways of the region are the Macal River, which is a branch of the Belize River to the north, and the Chiquibul River, which joins the Mopan (or

Western Branch) in Guatemala. There are also many small, primarily seasonal, streams that cannot be relied on as year-round water sources (Healy et al. 1983:399-400). Cotton

Tree Creek is one such water source that divides Caledonia in two. Awe (1985:31) suggests that both Cotton Tree Creek and the Macal River would have supported a variety of aquatic life which would have supplemented the subsistence resources at the site.

Similarly, a currently dry creek bed located along the southern base of the hill on which 8

Plazas A and B are situated may have been flowing when the site was occupied (Awe

1985:31) and may have provided additional aquatic and amphibious food resources.

The soils that characterize the eastern side of the Macal River comprise granite, shale, sandstone, quartzite and some limestone. The acidity caused by these materials has resulted in soil that is unsuitable for arable farming. The area does, however, support what Wright and colleagues (1959, cited in Awe 1985:16) call "orchard savannah", characterized by deciduous seasonal forests which include a mixture of oak (Quercus citrifolia; Q. hondurensis; Q. sapotaefolia), pine {Curatella americana; Clusia sp.), silver pimento palms (Thrinax argentea) and craboo {Brysonima crassifolia). The site of

Caledonia, however, is located within the rainforest and not the Pine Ridge, and thus the soils are nearly entirely derived from limestone bedrock (P. Healy, personal communication, 2010). With the assistance of terraced fields surrounding the site, which have been interpreted as intentional modification of the environment in order to increase food production (Healy et al. 1980), it is probable that the Caledonia Maya were able to support themselves sufficiently by agriculture. Awe (1985:32) argues that the close proximity of these features to the site's periphery demonstrates their association with

Caledonia's agricultural and subsistence base.

The nearby forest of the Pine Ridge region would have provided a wide variety of resources. As it is possible that pine was an important commodity for basic ritual practices (Morehart et al. 2005), Caledonia would have greatly benefited from the export of pine and pine resin for torches and pitch (Awe 1985:32). In addition, the site was located near one of the only known sources of granite (used for making manos and metates) in the Maya Lowlands, and it is possible that the residents of Caledonia were trading granite river cobbles throughout the region (Awe 1985:3). The presence of manos and metates within the site (Awe 1985:32) suggests that the residents of Caledonia were processing plants locally for consumption.

2.2 Archaeological Investigations and Burial Descriptions

The site of Caledonia (Belize Department of Archaeology Site No. 28/186-3) was a small ancient Maya centre of primarily Late Classic period date, although the site was occupied as early as the Late Preclassic or Proto-Classic period (A.D. 100-350) to the

Terminal Classic and perhaps into the Early Postclassic (A.D. 900-1000) periods (Awe

1985:388; see Table 2.1). The site was discovered in 1980 when reports of looting reached Dr. Paul F. Healy, who was searching for evidence of ancient Maya agricultural terraces and settlement around nearby Zayden Creek and . The site was brought to the researcher's attention by members of the Scottish Highland "Black Watch" Battalion of the British Army, who were conducting manoeuvres in the region at the time. It was . members of the Black Watch who suggested that the site be named Caledonia, and the modifier (Cayo) was added to distinguish the site from another called Caledonia located

Table 2.1: Time Periods Used at Caledonia

Maya Long Count Time Span Time Period Phase (Smith 1955:106) (Awe 1985:128)

Late Preclassic Chicanel ? to 8.12.0.0.0 300 B.C. to A.D. 250

Tzakol 1 8.12.0.0.0 to 8.16.0.0.0 A.D. 250 to 325 Early Classic Tzakol 2 8.16.0.0.0 to 9.1.0.0.0 A.D. 325 to 500 Tzakol 3 9.1.0.0.0 to 9.8.0.0.0 A.D. 500 to 600 Tepeu 1 9.8.0.0.0 to 9.13.0.0.0 A.D. 600 to 675 Late Classic Tepeu 2 9.13.0.0.0 to 10.0.0.0.0 A.D. 675 to 800 Tepeu 3 10.0.0.0.0 to 10.3.0.0.0 A.D. 800 to 900 Postclassic - - A.D. 900/1100 to 1500* •Date from Sharer and Traxler (2006:156). 10 in northern Belize.

Caledonia was investigated in 1980 by the Trent-Cayo Archaeological Project funded by a research grant (# 410-78-0373) from the Social Science and Humanities

Research Council (SSHRC) of Canada issued to Dr. Paul Healy, who directed the project.

This initial investigation focused on Plazas A and B. In 1984, further investigations at the site were carried out by the Trent University Field School, funded in part by Trent

University and the SSHRC Research Committee. Originally the second round of investigations were going to focus on specific structures within Plazas A and B, but shifted to include Plazas C and D upon their discovery.

The site consisted of two mound clusters (Plazas A and B, and Plazas C and D) located on a narrow strip of land along the southwestern bank of the Macal River and covered an area of approximately two hectares (five acres). In the northwestern end of the site, two connected mound clusters, Plazas A and B, were located on a hilltop and Plazas

C and D were located on a flattened ridge roughly 31m below and 120 m to the southeast. Plaza C is the largest at the site, and an open-ended ballcourt was built nearby.

No carved monuments have been identified at the site, although an elaborate stucco frieze was constructed on Str. C-l between the Terminal Classic and Early Postclassic periods.

A masonry vaulted temple was constructed atop Str. A-1 during the Tepeu 2 and 3 phases, and another atop Str. C-l with the addition of two benches with corbelled arch- shaped niches constructed during the Terminal Classic to Early Postclassic periods (Awe

1985:94-98). The discussion below focuses oh the areas where burials were recovered, and a full discussion of Caledonia can be found in Awe (1985).

Because of limited time and resources, only a fraction of the structures at the site 11 were investigated. Therefore, burials were only encountered in Structures A-1, C-l and

C-2, the most prominent buildings at the site. The burials have been described in detail elsewhere (Awe 1985) and are summarized below. A detailed description of the individuals analyzed from these burials is provided in Chapter Six. Here, the definitions of the terms "burial" and "grave" follow those provided by Smith (1972a:212), and were employed by the original researcher to describe the collection (Awe 1985: 101). Thus, a grave is the receptacle for a burial, and a burial includes anything connected with an interment, including the grave, skeletal material, and associated objects. Skeletons within a grave have been referred to as "Individuals", (i.e., Individual A) following Helmuth

(1985). Burials are further subdivided into four types: urn, simple, cist and tomb burials.

A simple burial consists of either an unlined hole in the ground or the inclusion of a body in the fill during the construction of a building. Any grave with definite borders, either stone walls or the sides of an excavation into structural fill, and lacking a capstone is classed as a cist burial. A grave that is more carefully walled, more elaborate, and covered with either capstones or wooden beams is considered to be a crypt or tomb. Both crypt and tomb have been used interchangeably to classify the same burials in the original report (Awe 1985). In this study the term tomb is adopted, although those at Caledonia are not as elaborate as some royal tombs found at larger Maya sites). Finally, an urn burial consists of the remains of an individual interred within a ceramic vessel. Table 2.2 provides a summary of the information discussed.

2.2.1 Plaza A - Burials #1 and #2

Plaza A was roughly one metre higher than Plaza B and consisted of six mounds.

Structure (henceforth Str.) A-1 was the largest in this Plaza, and showed evidence of Table 2.2: Summary of Burials from Caledonia* Burial Burial Burial Number of Time Period Age and Sex Number Type Location Individuals 1 Tomb Str. A-1 Early to Late Classic Tzakol 3 to Tepeu 2 Q** 1 Subadult, 8 Adults 2 Urn Str. A-1 Late Classic Tepeu 2 1 1 Subadult 3 Adult Males, 3 Cist Str. C-2 Late Classic Tepeu 2 4 1 Adult Female 2 Older Adult Males, 4 Tomb Str. C-2 Late Classic Tepeu 1 to 2 3 1 Middle Adult Female 5 Simple Str. C-l Late Preclassic Chicanel 1 1 Older Adult Female *As described by Awe (1985) and Helmuth (1985) ** Number of individuals identified by Healy and colleagues (1998) 13 being looted in the past. Burials #1 and #2 were located within this structure and are discussed below.

2.2.1.1 Burial #1

During the 1980 excavation, investigators came across a large, rectangular-shaped capstone of a tomb 65 cm below the surface (Awe 1985:43). The oval chamber below was vaulted and was categorized as a tomb. This tomb has been interpreted as a multiple burial, where individuals were interred in succession, as the chamber was not large enough to fit more than one fleshed body at a time (Healy et al. 1998). Nine poorly preserved individuals were reported from this burial, at least one of which was a subadult

(see Table 2.2). Although the fragmentary nature of the remains made the position of the different individuals difficult to distinguish, at least one individual was fully extended

(Awe 1985; Helmuth 1985; Healy et al. 1998).

Skeletal remains were oriented east-west, with the crania facing west. Burial offerings included 17 pottery vessels, one land crab (Gecarcinidae) claw, 32 disk-shaped shell (possibly Spondylus sp.) beads, two conch (Strombus sp.) shell buttons, two jade . pendants, four jade beads, one spindle whorl, one mano fragment, 10 mussel (probably

Nephronaias sp.) shells, and two obsidian blade fragments (Awe 1985:103). The ceramics indicate that the tomb was in use between the late Early Classic (Tzakol 3) and Late

Classic (Tepeu 2) periods, and a sample of charcoal from the burial was radiocarbon dated to 1520 ± 120 years B.P. (A.D. 460 ± 140 years MASCA Correction) (Awe

1985:103,430).

2.2.1.2 Burial #2

Along the primary axis of Str. A-1 where the platform joined the stairway a large 14 broken bowl was recovered. It was lying on its side with its orifice facing south and its base facing north towards Plaza A. Two teeth and four fragments from the long bones of a human subadult were found within the vessel (Awe 1985:49). The pottery vessel dated to the Late Classic (Tepeu 2) period. This was classed as an urn burial, defined as the remains of an individual within a pottery vessel (typically a large jar) sometimes capped by a lid or inverted dish or plate (Andrews and Andrews 1980:314).

2.2.2 Plaza C - Burials #3, #4 and #5

Plaza C comprises a large open patio enclosed by four mounds (Str. C-l through

C-4), a ballcourt (Str. C-7 and C-8), and what have been interpreted as residential structures (Str. C-5, C-6 and C-9). Structure C-l is the most impressive structure at the site, measuring five by four meters at its base and five metres tall (Awe 1985:27). The earliest burial at the site (Burial #5) was recovered from within this structure. A looter's trench was also discovered on the top of the mound. Structure C-2 is the second tallest structure at the site and contained Burials #3 and #4. It was determined that Plaza C was the central precinct of the site and, because of the presence of a looter's trench on Str. C-l and the chronology that could be provided from Str. C-l and C-2, this plaza became the focus of the 1984 archaeological investigations (Awe 1985:33-34).

2.2.2.1 Burial #3

Burial #3, a cist burial, was uncovered in a natural soil level below Plaza C near

Str. C-2: It was located 42 cm below the destroyed section of the plaza floor, running along the primary axis of the structure, extending westward from the base of Stairway 1.

Due to the location of this burial, it is most likely intrusive (Awe 1985:69). Burial furniture included seven miniature eccentrics (elaborately carved stone objects), a 15

tapered-stemmed point, an obsidian core, and what is likely a ballcourt marker over the

skull of an extended individual. Ceramics recovered from the burial suggest that it was

deposited late during the Late Classic (Tepeu 3) (Awe 1985:69). Four adults, three probable males and one probable female, between the ages of 20 and 40, were recovered

from this burial (Awe 1985:110; Helmuth 1985:418-422). During recovery it was found that two of the individuals were interred in a sitting position (Awe 1985:68), although it is

not clear today which individuals these were.

2.2.2.2 Burial #4

A second tomb, Burial #4, was located 10 cm below the first step of the stairway

of Str. C-2. The rectangular chamber lay perpendicular to the primary axis of the structure

in a north-south alignment and had been filled with sandy river soil, likely due to

"practical architectural considerations" (Awe 1985:71). It is likely that this burial was

intrusive during the second construction phase of Str. C-2 (Awe 1985:99).

Two individuals recovered from this burial were identified as male adults at least

40 years of age at death, and a third was identified as a probable female between the ages

of 20 and 30 years at death (Awe 1985:113; Helmuth 1985). At least one of the

individuals was laid in an extended position with their face up, and it is likely that the

other two individuals were placed this way as well. Burial furniture included 12 ceramic vessels, one jade bead, three jade or jadeite ear ornaments, one shell bead, two bone needle fragments, two antler fragments, six obsidian blade fragments, two spindle whorls

and four fragments of a shell mosaic (Awe 1985:113). The tomb was dated to the Late

Classic (Tepeu 1 and 2) period. Awe (1985:114-115) argued that the crypt was too small to accommodate all three individuals at once and, instead, suggested that two of the 16 individuals represented sacrificial victims who were to accompany the third individual, a noble personage, to the afterlife.

2.2.2.3 Burial #5

Eighty-five centimetres below floor 7 in Str. C-l, a simple burial containing the remains of a single, older adult female was unearthed. This individual was placed in a flexed position and a single body sherd, possibly Late Preclassic in date, was associated with the burial (Awe 1985:84). This individual almost certainly represents one of the earliest inhabitants of Caledonia, and the lack of grave goods could either indicate a male bias in society, or, more likely, "the relative poverty of the site's population in the Late

Preclassic" (Awe 1985:116). This burial probably represents the pattern of simple, flexed, single burials that was predominant throughout the lowland Maya region in the Late

Preclassic and Early Classic periods (Awe 1985:116).

2.2.3 Additional Artefactual and Faunal Remains

In addition to the skeletons described, burial artefacts such as non-human bone, shell, and ceramics were identified. These are listed in Appendix A, and detailed descriptions of the archaeological investigations at Caledonia can be found in elsewhere

(Awe 1985). 17

CHAPTER THREE

THEORETICAL PRINCIPLES OF STABLE ISOTOPE ANALYSIS

3.1 The Principles of Stable Isotopes

Elements are distinguished by the number of protons in their nucleus; carbon has

six, and nitrogen has seven. Isotopes are composites of the same element with differing numbers of neutrons in their nuclei. Carbon has two stable isotopes: carbon-12 ( C), which is relatively abundant, and carbon-13 (13C) which is less common in nature.

Similarly, the two stable isotopes of nitrogen are nitrogen-14 (14N), which is relatively abundant, and nitrogen-15 (15N), which is less common. The isotopes of an element possess nearly identical chemical properties but have different atomic masses. This difference in mass affects how isotopes are incorporated into the tissue of an organism because the heavier isotope is slower to respond in a chemical reaction (Hoefs 2004;

Peterson and Fry 1987; Urey 1947).

The relative abundances of stable carbon or nitrogen isotopes are calculated by measuring the ratio of 13C to 12C or 15N to 14N relative to the isotope ratio of a standard reference material. The isotopic composition of an element is expressed using the delta

(8) notation, and the unit of measurement is per mil (%o) (Figure 3.1). The standard reference material for carbon is the International Atomic Energy Agency (IAEA) standard material of NBS-19 which has a 813C value of+1.95%o relative to the exhausted peedee belemnite (VPDB) standard (Coplen 1994). The standard reference material for nitrogen is the atmospheric N2 ambient inhalable reservoir (AIR) standard (Mariotti

1984).

The majority of fractionation discriminates against the heavier stable isotopes 13C 18

13^/12^ 13/—/12/- 13/-1 _ \^l ^sample— «-" ^standard 5 — M%o) Op/lip, l-V ^-standard

15XT/14XT 15-KT/14> 15 5 N %„l = N/ NSampie- N/ Nstandard

N/ Nstandard

Figure 3.1: The isotopic composition of an element. or 15N which produces a lower 8 value where the product is depleted of the heavier isotope. However, if the fractionation discriminates against the lighter 12C or 14N stable isotopes, the 8 value will be higher and the product is enriched in 13C or 15N. The VPDB standard for carbon contains more C than nearly all dietary resources and human tissue

(Ambrose 1993:65). Therefore the 813C values for carbon in most biological systems are depleted of C relative to the VPDB standard, and have negative values (i.e., < 0%o).

Alternatively, the 815N values in most biological systems are enriched relative to the standard (AIR), and thus have positive values (i.e., > 0%o).

It is well established that the isotopic composition of a consumer's tissues is an accurate reflection of their diet (Ambrose and Norr 1993; DeNiro and Epstein 1978,

1981; Hare 1980; Macko et al. 1982; Tieszen and Fagre 1993). Once the isotopic composition of local foods (i.e., the menu) is known, it is possible to determine, based on comparisons, the actual meal. Although stable isotope analysis cannot precisely measure the quantity of each food item consumed, this method does provide direct evidence for:

(1) the types of plant foods consumed (C3/C4 plants, legumes vs. non-legumes); (2) the relative contribution of protein to the diet; (3) the types of animals consumed

(marine/terrestrial, omnivores/herbivores/carnivores); and (4) variability in the diet

(mono-isotopic vs. multi-isotopic diets). 19

3.2 Stable Carbon Isotopes

3.2.1 Carbon and Dietary Reconstruction

Reviews of the application of stable isotopes to investigate archaeological diet have been provided elsewhere (Ambrose 1993; Bumsted 1985; DeNiro 1987; Katzenberg

1988, 2008; Pate 1994; Pollard and Heron 2008; Price et al. 1985; Schoeninger 1989;

Schoeninger and Moore 1992; Tykot 2006) and are briefly summarized here. The use of carbon isotopes for dietary reconstruction was recognized in the early 1960s when Parker

(1964) applied stable carbon isotope analysis to investigate isotopic differences between marine plants and animals. In the following decades, van der Merwe and Vogel used carbon isotope analysis to investigate the diet of a human skeleton from Transvaal, South

Africa (van der Merwe 1982), and demonstrated the introduction of maize agriculture in

North America (van der Merwe and Vogel 1978; Vogel and van der Merwe 1977) and

Venezuela (van der Merwe et al. 1981). Subsequent studies of stable carbon isotope analysis have investigated dietary trends worldwide, including those of the ancient Maya

(to be discussed in greater detail in Section 4.3).

3.2.2 Carbon in Terrestrial Ecosystems

Carbon dioxide (CO2) in the atmosphere contains roughly 99% of the stable isotope C, and the majority of the remaining 1% consists of the stable isotope C (Faure

1986:491; Smith 1972b). During photosynthesis, plants discriminate against the heavier

13C isotope because of minute differences in chemical and physical properties caused by the difference in mass (see O'Leary 1988). Plants that discriminate most against C have

1 "\ lower, or more negative, 8 C values compared to the carbon standard (average of

-26.5%o; O'Leary 1988; Smith 1972b; Smith and Epstein 1971). These are known as C3 20 plants (Calvin-Benson photosynthetic pathway), because they fix carbon from ribulose diphosphate into a three-carbon molecule using the enzyme ribulose bisphosphate carboxylase (Calvin and Benson 1948). These plants include the majority of wild plants in Mesoamerica, most vegetable cultigens, nuts, fruits and many temperate grains including wheat, barley and rice.

Plants that incorporate more 13C during photosynthesis are called C4 plants

(Hatch-Slack photosynthetic pathway) because they fix CO2 into phosphoenolpyruvate to yield four-carbon acids using the enzyme phosphoenolpyruvate carboxylase (Hatch and

Slack 1966; Hatch et al. 1967). These plants have higher, or more positive, 8 C values compared to C3 plants (average of -12.5%o; O'Leary 1988; Smith 1972b; Smith and

Epstein 1971), and include tropical grasses such as sorghum, millet, and maize, the major cultigen in Mesoamerica. Archaeological maize from southern Ontario, Canada, however, has been found to have an average 813C value of -9.5%o (Schwarcz et al. 1985). Because the isotopic signatures of C3 and C4 plants do not overlap (there is a 3 to 12%o difference between the two worldwide; Ambrose 1993; Tieszen and Boutton 1989) diets based on either plant type can be inferred.

A third pathway, known as crassulacean acid (CAM), occurs in succulent plants, which utilize both C3 and C4 photosynthetic pathways, depending on the environment. As a result, CAM plants have a wide range of 813C values (-27 to

-12%o) which overlap the values for C3 and C4, thus complicating interpretations of past diet. In the Southern Maya Lowlands, CAM plants tend to overlap with C4 plants (see

Wright 1997a: 187). One CAM plant of concern in this region is the agave (Agave spp.), which was cultivated throughout Central America and currently grows in the Maya

Highlands and Northern Lowlands. Early Mesoamericans discovered that if the heart of 21 the plant at the base of the stalk is removed before it flowers, the plant releases aguamiel into the cavity. This substance can be produced for months, making it a valuable food resource during times of food shortages (Crist 1939). At the site of Ceren, El Salvador, it appears that Classic Maya commoners in the periphery of the Highlands were growing agave for production of fibres for making artefacts (Lentz et al. 1996; Sheets 2000).

While there is no evidence to suggest that the Maya at Ceren were not also cultivating agave for consumption, it is unlikely that the Caledonia Maya consumed this plant (see

Section 4.2).

Other plants that may have been available to the Maya that could affect the stable carbon isotope results include the C4 plants amaranth and epazote, and several CAM plants including yucca, pineapple, pinnuela, prickly pear and nopal cactus (Coyston

1995:84; Wright and White 1996). It has been suggested that these resources were only consumed in times of need (i.e., famine periods) and thus it is unlikely they were consumed in any great quantity by the ancient Maya (Powis et al. 1999; White 2004:362).

Thus, the most likely plant represented by a C4 signature in the isotopic composition of human skeletons from Caledonia is maize (Hellmuth 1977; Tozzer 1941; Turner and

Harrison 1981).

3.2.3 Carbon in Marine Ecosystems

In marine environments carbon is derived from several sources (Hoefs 2004). The first source of carbon is dissolved bicarbonate (HC03_) and carbonic acid (H2CO3), which are enriched in I3C relative to atmospheric CO2 used by terrestrial plants (81 C

1 3 value of-7%o), and thus have 8 C values of ~0%o (Smith and Epstein 1971). Marine plants also obtain carbon from terrestrial detritus washed into oceans by rivers, which 22 contains a mix of S13C values from local terrestrial plants, as well as dissolved CO2, with

8 C values of atmospheric CO2 (-7%o). Marine food chains are based primarily on C3 plants; however, because they utilize carbon from dissolved bicarbonate rather than atmospheric CO2, their values are enriched relative to terrestrial C3 plants and have a value of roughly -19.5%o (Smith 1972b; Smith and Epstein 1971). The exceptions to this are coastal and estuarine ecosystems, where sea grasses with 8 C values of roughly

-13%o contribute to the food web (Chisholm et al. 1982; Keegan and DeNiro 1988;

Schoeninger and DeNiro 1984). If a marine signature is found at Caledonia, the resources likely would have originated from the sea grass-based system off the coast of Belize in the Caribbean Sea (see Keegan and DeNiro 1988). It can be difficult to distinguish between marine and maize resources in the diet of individuals who lived in regions where both resources were available and consumed, which may be the case at Caledonia, because their isotope ratios overlap (Ambrose 1986).

3.2:4 Carbon Isotopes in Freshwater Ecosystems

Freshwater plants (i.e., those in lakes, rivers and streams) also use dissolved CO2 as a carbon source, although in general the 813C values are more negative than those found in marine ecosystems. As a result, freshwater plants and animals have more

1 ^ negative 8 C values than marine plants and animals. Plant detritus, consisting primarily of C3 plants in the Maya region, provides a second source of carbon utilized by coastal marine and some freshwater plants.

3.2.5 Complicating Factors that Influence 813C Values

The isotopic composition of C4 plants appears to be relatively unaffected by environmental factors such as humidity, light intensity and temperature (O'Leary 23

1 3

1988:334). While the variations in 8 C values of C4 plants are not well understood, there are several environmental factors that influence the carbon isotope ratios of C3 plants.

These include water availability, light intensity, temperature, partial of CO2 and nutrient availability (Smith and Epstein 1971; Tieszen 1991; Ambrose 1993). For example, unpublished data from O'Leary and Treichel (1987; see O'Leary 1988) found that 813C values of C3 plants become more negative with increasing temperature, and became more positive with increased light intensity. 1 ^

Fossil fuels are depleted of C; as such, the burning of fossil fuels has decreased the overall 8 C of atmospheric CO2. Modern plants have 8 C values that are 1.5%o lower than pre-industrial plants (Friedli et al. 1986; Keeling et al. 1979; Marino and McElroy

1991), and in prehistory the 813C values of C3 and C4 plants were probably closer to

-25%o and -1 l%o, respectively (DeNiro 1987). Therefore, when modern plants and animals are sampled to represent their ancient counterparts in dietary reconstruction, it is necessary to add 1.5%o to their values (see Section 4.3.1).

The understories of forests have distinctive carbon isotope compositions because biogenic CO2 is recycled. The decomposition and of C3 plants in these areas produces CO2 with a value similar to the source flora, and the canopy prevents rapid mixing of atmospheric and biogenic CO2 (Jackson et al. 1993; van der Merwe and

Medina 1989, 1991). This is called the "canopy effect", which lowers the 813C of atmospheric CO2 near the forest floor, causing leaves in this zone to have lower (i.e., more negative) S13C values than those higher in the canopy. As it is likely that humans, and the animals they consumed, would rely on vegetation near the forest floor, the more negative 813C values of such plants resulting from the canopy effect will be passed onto their consumers. 24

The isotopic variation in CAM plants is also a function of environmental

conditions (i.e., day length, salt salinity, water stress, and night temperature). For

example, in hot, dry areas CAM plants exhibit 8 C values similar to C4 plants, while in

cooler environments they more closely resemble that of C3 plants (Troughton et al.

191 A). In areas where both C4 and CAM plants are available, they cannot be differentiated with stable carbon isotope analysis (Ambrose 1993:92). As discussed, the ancient Maya did not likely consume CAM plants in great quantity and thus this should not affect the results of this study.

Finally, in areas where marine resources have stable carbon isotope ratios similar to those of terrestrial C4 plants it is not possible using carbon isotope data alone to determine the contribution of terrestrial and marine resources in the diet because of isotopic overlap (Chisholm et al. 1983; Schoeninger and Moore 1992; Walker and

DeNiro 1986; White et al. 1993). In such a situation, it is essential to combine stable carbon isotope analysis with stable nitrogen isotope analysis, and to assess the collagen-

13 to-bioapatite carbon spacing (A Cbi0-coi) to distinguish marine from terrestrial resources.

3.3 Stable Nitrogen Isotopes

3.3.1 Stable Nitrogen Isotopes and Dietary Reconstruction

The analysis of stable nitrogen isotopes for dietary reconstruction was developed in the 1980s. This was driven by indications that dietary differences in stable nitrogen isotope values were caused by trophic level effects, particularly in marine ecosystems

(Ambrose and DeNiro 1986; DeNiro and Epstein 1981; Schoeninger and DeNiro 1983,

1984; Schwarcz et al. 1985; Sealy et al. 1987). Additional studies have since increased 25

our understanding of nitrogen isotope ratios, specifically regarding the effects of climate

and environment on values, as well as trophic level increases in terrestrial and marine

ecosystems, which will be discussed in greater detail below.

3.3.2 Nitrogen in Terrestrial Ecosystems

Nitrogen in the atmosphere is composed of roughly 99.6% 14N and 0.4% 15N, and

is present primarily as N2 gas (S15N = ~0%o), but also as nitrous oxide (N2O) and nitrogen

dioxide (NO2) (Faure 1986:513-514). Microorganisms facilitate the conversion of

nitrogen in the three stages of its biological cycle (fixation, nitrification, and

denitrification) with varying degrees of fractionation (Hoefs 2004:49-51). Terrestrial

plants can be divided into two groups based on their stable nitrogen isotope ratios:

legumes and non-legumes (DeNiro 1987). Legumes (i.e., beans, peas and peanuts) fix

atmospheric N2 and have 815N values close to 0%o (Delwiche and Steyn 1970; Wada et al.

1975). Non-legumes rely on nitrogen in the soil because they cannot fix N2, and have

815N values that average +3%o (Delwiche and Steyn 1970).

3.3.3 Nitrogen in Aquatic Ecosystems

Most nitrogen in the oceans is present as dissolved N2 (815N = ~0%o), but

ammonia (NH3), nitrite (NO3 ) and nitrate (NO2 ) are also present in varying

(Faure 1986:513). In freshwater lakes and rivers, nitrogen is present as

ammonia, protein, nitrate and amino acids in (Faure 1986:514). Marine and terrestrial food chains have distinct 815N signatures (Schoeninger and DeNiro 1984; Sealy

and van der Merwe 1986; see Figure 3.5). Marine plants cannot fix N2 and instead fix

dissolved nitrate and ammonium, and have a wide range of S15N values with a mean of

+7%o (DeNiro 1987). On average, the 815N values of marine plants are roughly 4%o higher 26

than terrestrial plants (Delwiche and Steyn 1970). One exception is cyanobacteria (blue-

green algae) which fix atmospheric N2 dissolved in the ocean and have a 815N value of

~0%o (Delwiche and Steyn 1970; Wada et al. 1975). In tropical marine reef and mangrove

ecosystems, which were likely exploited by the ancient Maya, blue-green algae contribute

a significant amount of nitrogen to the food chain. As a result, these ecosystems will

exhibit lower 815N values than marine ecosystems that approach those of terrestrial

ecosystems (Schoeninger and DeNiro 1984; Schoeninger et al. 1983).

3.3.4 Complicating Factors that Influence 515N Values

Nitrogen isotope values can be affected by a variety of factors, including climate, physiology, pregnancy, and pathological conditions (Ambrose 1986, 1991, 2000;

Ambrose and DeNiro 1986; Aufderheide et al. 1988; Delwiche and Steyn 1970; DeNiro

and Epstein 1981; Fuller et al. 2004; Hare 1980; Heaton et al. 1986; Katzenberg and

Lovell 1999; Schoeninger et al. 1983; Schwarcz et al. 1999; Sealy et al. 1987; Wada et al. 1975; Waters-Rist and Katzenberg 2010; White and Armelagos 1997). For example,

laboratory experiments have demonstrated that contact with moisture and high can drastically influence the amount of nitrogen present in collagen (Hare

1980). Researchers have also determined that elevated 815N values in animals are associated with increased aridity (Ambrose 1991; Heaton et al. 1986). This, however, is related to the finding that water-stressed animals have 815N values 2%o to 4%o higher than non-stressed animals (Ambrose 1986), which represents physiological fractionation of nitrogen, rather than climatic influence. Furthermore, recent feeding experiments have found evidence that disputes the claim that climate impacts the composition of nitrogen in bone (Ambrose 2000). 27

Pathological and physiological conditions such as pregnancy can also influence

the nitrogen composition of bone (Katzenberg and Lovell 1999; White and Armelagos

1997). For example, elevated 815N values have been observed in association with

osteopenia rather than diet in Nubian mummies (White and Armelagos 1997). Similarly,

stable isotope analysis of nitrogen in human hair has indicated that the 815N values of women with adequate nutrition decreased by roughly l%o during pregnancy, which is referred to as the pregnancy effect (Fuller et al. 2004). An alternative effect wherein nutritionally stressed pregnant womenexperienced an increase in 815N values has also been reported (Fuller et al. 2005). The turnover rate of collagen, however, is much slower than that for hair keratin, reflecting an average of the last several years or decades of an

individual's life, and it is more difficult to detect the pregnancy effect in human bone

collagen compared to hair keratin (Fuller et al. 2006; Nitsch et al. 2010). According to a model proposed by Nitsch and colleagues (2010), unless a woman died soon after giving birth, the depletion in 815N values experienced as a result of pregnancy would not be

detectable either because of collagen turnover or because it would be masked by dietary

differences in S15N values. Finally, a recent study on a sample of fully weaned subadults

from a Protohistoric ossuary in southern Ontario has demonstrated that 815N values are not affected by growth during childhood (Waters-Rist and Katzenberg 2010).

Marine ecosystems based on non-nitrogen fixing plants will exhibit increased

815N values compared to their terrestrial counterparts. This is because marine food chains contain many more trophic levels (e.g., primary, secondary and tertiary carnivores) than terrestrial food chains which cause animals at the highest trophic level to be much more enriched in 15N relative to terrestrial food webs (Schoeninger and DeNiro 1984). The ancient Maya, however, likely exploited marine and reef fish of varying trophic levels, 28 which have average 815N values that may overlap with those of terrestrial resources

(White et al. 2001b). Finally, in the Maya area, the 815N values of freshwater fish are elevated are enriched relative to terrestrial animals and marine resources (Wright 1994).

In contrast, freshwater molluscs and snails have 815N values that overlap those of terrestrial animals (Schoeninger et al. 1983). However, because there is substantial variation in 15N enrichment by trophic level in all ecosystems (Vander Zaden et al. 1997;

Vanderklift and Posnard 2003), it is important to interpret trophic level and protein

15 13 consumption using both N and collagen-to-bioapatite spacing (A Cbi0-coi) values.

3.4 Bone Biology and Diet

Bone consists of inorganic calcium phosphates precipitated in an organic collagen matrix. Cortical bone is comprised principally of an inorganic mineral phase (-70%) and water (9%) bound to an organic matrix (21%) (Triffitt 1980). Bone mineral is referred to as bioapatite and consists of a calcium phosphate [Cai0(PO4)6(CO3)(OH)2] that is similar to poorly crystalline hydroxyapatite [Cas(P04)30H] in its organic components and crystalline structure (Glimcher et al. 1981). This phase of bone also has high lability (the ease with which a substance can dissolve or undergo a phase change; Glimcher et al.

1981), increasing the likelihood of diagenetic alterations over time (see Section 3.6).

Bioapatite contains a small percentage of carbon present as carbonate (CO3 ') that occurs as absorbed ions on crystal surfaces or as structural carbonate which is substituted

3- into the crystal lattice at either the OH or PO4 positions. Absorbed carbonate is not associated with a well-defined lattice position and is instead absorbed either onto the surface of apatite crystal or incorporated into less mature regions of apatite crystals.

Because absorbed carbonate is more easily dissolved and altered than structural 29 carbonate, the latter will be less affected by diagenesis than the former (Krueger 1991;

Lee-Thorp 2000:91-92). Tooth enamel bioapatite has a higher crystallinity than bone bioapatite, and as such is less likely to be affected by diagenesis (Lee-Thorp and van der

Merwe 1987, 1991; Wang and Cerling 1994; Zazzo et al. 2004).

The protein collagen makes up approximately 90% of the organic portion of cortical bone. Collagen is essentially bundles of fibrous protein consisting of four amino acids (glycine, proline, hydroxyproline and alanine) (Krueger and Sullivan 1984:209), and contains roughly 45% carbon (Elliot 1994). The remaining 10% of the organic phase of bone is made up of noncollagenous proteins (including phosphoproteins, glycoproteins,

GLA proteins and proteoglycans), , carbohydrates, enzymes and hormones. Amino acids are the basic structural components of collagen fibrils, and the collagen molecule consists of three polypeptide chains coiled around each other to form a triple helix.

3.4.1 Bone Remodelling and Turnover Rates

The main purpose of bone remodelling and turnover is to maintain the load- bearing capacity of bone. Bone remodelling is carried out by temporary anatomic structures know as basic multicellular units (BMUs). Two cells within the BMUs are essential for bone maintenance: those involved in the dynamics of bone formation

(osteoblasts), and those involved in bone destruction or resorption (osteoclasts). During childhood, healthy bone primarily undergoes growth (the effect of osteoblasts), whereas into adulthood growth ceases and bone maintenance begins. Bone metabolism in healthy adults is thus characterized by the actions of osteoclasts and osteoblasts. The rate at which bone is removed and replaced (i.e., turnover) varies, and the chemical composition of 30 normal adult human bone will reflect long-term dietary averages as a result of the slow remodelling process.

It has often been assumed that dense cortical bone, such as that found in the diaphyses of long bones, remodels slowly and thus contains tissue deposited over many years, whereas cancellous bone is likely to turn over much faster, thus reflecting the isotopic composition of diet closer to the time of death (Hedges et al. 2007; Sealy et al.

1995). This assumption is based on the fact that remodelling occurs on the surface of bones, and cancellous bone has a high surface-to-volume ratio bone relative to cortical bone (Parfitt 2002). It appears, however, that the location of a bone, that is whether it is part of the central (i.e., axial skeleton including the pelvis) or peripheral (i.e., appendicular skeleton without the pelvis) skeleton, has a greater effect on metabolic functions than the type of bone (Parfitt 2001).

Estimates of the rate of bone turnover vary considerably, but bones are assumed to fully turn over every 10 to 20 years. For example, using radiocarbon as a tracer, Libby and colleagues (1964) demonstrated that collagen turnover in adults exceeded a decade in their sample. Similarly, Parfitt (2001) proposed that a low turnover rate of 2-5% each year

(i.e., complete turnover every 20 years or more) would be sufficient to maintain the mechanical competence of bone.

The varying rates that have been reported for bone turnover are also likely related to physiology and disease processes. For example, collagen turnover rates during adolescence (Hedges et al. 2007; but see Waters-Rist and Katzenberg 2010) and in menopausal women (Delmas 1993) have been found to occur more quickly, and rates in patients with untreated vertebral osteoporosis varied considerably (Delmas 1993). It is also difficult to distinguish between dietary, genetic and biomechanical influences on 31

cortical bone remodelling (see Pfeiffer and King 1983). Although the rate of bone

turnover varies depending on the sampled bone, all will reflect the diet of roughly the last

decade or so of life. This is a much longer period than is represented by soft tissue such as

hair keratin (Hedges et al. 2007).

In contrast, carbon in tooth enamel bioapatite is deposited at the time the tooth

was formed during childhood. Because enamel is not subsequently remodelled, the

carbon in this tissue reflects the dietary average of the period when the tooth was formed

early in life. In a sense, this tissue maintains an isotopic record from childhood, which can

be compared to the isotopic signatures in bone collagen and bioapatite or other teeth to

investigate changes in diet (using stable carbon isotopes; see Wright and Schwarcz 1998)

or mobility (using stable oxygen and strontium isotopes; see Price et al. 2010; Wright et

al. 2010) over the lifespan of a single individual.

3.4.2 Dietary Fractionation of Carbon and Nitrogen

1 "^

Carbon from plants is enriched in C as it is incorporated into the tissues of

13 higher order consumers (A Cdiet-c0i)5 but the magnitude of this fractionation is not

consistent and the reason for this is not well understood (Ambrose 1993:104). It has been proposed that the collagen of primary consumers is enriched in 13C by approximately 5%o relative to diet in both marine and terrestrial food chains (van der Merwe 1982; van der

Merwe and Vogel 1978). While it is now recognized that collagen enrichment is more variable (Vogel 1993; Bocherens and Drucker 2003), the 5%o spacing appears to hold true

in cases of maize dependency (Gerry and Krueger 1997). Controlled diet experiments have estimated a mean difference between diet and bioapatite 813C values (A13Cdiet-bio) of

9.5%o (Ambrose and Norr 1993; DeNiro and Epstein 1978; Tieszen and Fagre 1993). 32

Although there is a l%o enrichment of carbon at each trophic level (position in the food chain), it is only discernible in the best-controlled systems (Schoeninger and Moore

1992:258) and will not be discussed further here.

The trophic level effect is perhaps better associated with stable nitrogen isotope analysis. This is because a stepwise enrichment in 15N on the order of 3%o to 4%o in collagen is well documented between trophic levels in terrestrial as well as marine environments (Ambrose et al. 1997; DeNiro and Epstein 1981; Schoeninger and DeNiro

1984). Furthermore, this enrichment in 15N by trophic level also occurs in breastfeeding mammalian infants. In humans, breastfeeding infants have S15N values that are roughly

2%o higher than that of their mothers (Fogel et al. 1989).

3.4.3 Distribution of Carbon from Macronutrients to Collagen and Bioapatite

Food is composed of three macronutrients - carbohydrates, proteins and lipids - which are used by the body for tissue maintenance and growth. Initially, researchers assumed that all dietary macronutrients contributed equal amounts of carbon to bone collagen (Schoeninger 1989; Schwarcz 1991). With time, however, researchers began to question the validity of this assumption, arguing that collagen preferentially represents dietary protein, whereas bioapatite represents dietary energy (Krueger and Sullivan 1984;

Lee-Thorp et al. 1989). This is because the essential amino acids found in collagen cannot be synthesized by the body and must be obtained from the diet. Subsequent feeding experiments have demonstrated that dietary protein is preferentially routed to bone collagen (Ambrose and Norr 1993; Howland et al. 2003; Jim et al. 2004; Tieszen and

Fagre 1993).

The analysis of stable carbon isotopes in bone collagen will reflect the isotopic 33 composition of the whole diet in circumstances where the diet is vegetarian, and in areas where the carbon isotope ratios of plants and animals are the same (i.e., mono-isotopic).

However, in areas where protein and non-protein resources have different carbon isotope ratios (i.e., multi-isotopic), collagen may not accurately reflect the isotopic composition of the whole diet (Ambrose et al. 1997). Although the ancient Maya utilized a variety of floral and faunal resources, it is assumed that their dietary staple was maize, a C4 plant

(see Chapter Four). If the Maya consumed terrestrial animals or freshwater fish feeding on C3 plants, their collagen isotope values would over-represent a C3 signature and under-represent the C4 plant component of their diet. Alternatively, if protein was primarily derived from marine resources, the carbon isotope ratio in collagen would incorrectly reflect a maize-based diet.

In contrast, carbonate in bone bioapatite is formed by the precipitation of blood bicarbonate and unused macronutrients are respired as CO2 from the lungs. Thus, carbon isotope ratios in bone bioapatite will reflect carbon from carbohydrates and lipids, and the amount of carbon from protein will depend on the amount of protein in the diet (Ambrose

1993; Howland et al. 2003; Jim et al. 2004; Tieszen and Fagre 1993; but see Kellner and

Schoeninger 2007). The isotopic analysis of bone bioapatite can therefore mitigate the problems associated with the interpretation of 8 C values from collagen.

Following this, researchers proposed that the offset between 8 Ccoi and 8 Cbio

13 (i.e., A Cbi0-coi) was also related to diet, and varied depending on whether the consumer was herbivorous, omnivorous or carnivorous (Krueger and Sullivan 1984; Lee-Thorp et al. 1989). Herbivores should demonstrate a large difference between collagen and bioapatite 8 C values (denoted as A Cbio-coi), because the difference in 8 C values between plant protein and carbohydrates is small. In carnivorous diets, lipids rather than 34 carbohydrates are depended on for a portion of energy metabolism. Because lipids are

1 ^ depleted of C relative to carbohydrates or proteins (Vogel 1978; DeNiro and Epstein

13 1978), carnivores exhibit smaller A Cbi0-coi values than herbivores. The spacing between 1 "^ collagen and bioapatite 8 C has thus been used as an index for meat consumption

(Krueger and Sullivan 1984; Lee-Thorp et al. 1989; but see Kellner and Schoeninger

2007). 1 ~\

The fluctuation in A C^o-coi was found to be largely affected by the percentage of protein in the diet and the carbon composition of the diet (Ambrose and Norr 1993:29). If the diet is inadequate in quantity and quality of protein, carbon from energy substrates

(primarily carbohydrates from plant foods rather than lipids) is routed to collagen during amino acid synthesis, as well as to bioapatite. In this case, the bioapatite-collagen spacing is larger and suggests an herbivorous diet (Krueger and Sullivan 1984).

An average spacing of roughly 7%o has been reported for herbivores, 5%o for omnivores and 3-4%o for carnivores (Ambrose 1993; Clementz et al. 2009; Krueger and

Sullivan 1984; Lee-Thorp et al. 1989), although it is possible these values reflect more than a dietary difference (Hedges 2003). Finally, marine-based diets have been found to create small collagen-to-bioapatite spacing values (i.e., < 3%o) that exaggerate the appearance of carnivory (Ambrose and Norr 1993; Lee-Thorp et al. 1989; Norr 1995:207;

White et al. 2001b). Such small values are useful in differentiating between marine and terrestrial carnivorous diets.

The A Cbio-coi is also useful for differentiating between mono-isotopic and multi- isotopic diets. When the protein and non-protein components of the diet have similar 813C values the difference between collagen and diet is +5%o, and because the difference between bioapatite and diet is always around 9.4%o, a mono-isotopic diet would exhibit a 35

13 13 A Cbio-coi of+4.4%o (Ambrose et al. 1997:351). When the 8 C values of dietary protein

1 "\ are more negative than that of the whole diet, A Cbio-coi is greater than 4.4%o, which would be the case if the diet consisted primarily of C4 carbohydrates and C3 proteins. If 1 ^

A Cbio-coi is less than 4.4%o, dietary protein is less negative than the whole diet which is consistent with a diet that is primarily comprised of C3 carbohydrates and marine proteins

(Ambrose et al. 1997:351).

3.5 Diagenesis

Before stable isotope data can be used to reconstruct diet, it is necessary to establish whether the isotopic compositions of the samples have been diagenically altered.

Diagenesis refers to the postmortem alteration of the chemical composition of bones and teeth following their deposition in the burial environment. As discussed earlier, the isotopic composition of a sample of bone or tooth from an individual has a direct relationship to the foods they consumed (Ambrose and Norr 1993; DeNiro and Epstein

1978, 1981; Howland et al. 2003; Krueger and Sullivan 1984; Schwarcz 1991; Tieszen et al 1983). If diagenesis occurs, the chemical composition of a bone no longer reflects the diet (i.e., biogenic), but instead the post-depositional environment (i.e., diagenic). As such, stable isotope values measured in bones and teeth may not accurately reflect diet.

Consequently, a number of studies have assessed the impact of diagenesis on dietary investigations using biogeochemistry (i.e., El-Kammar et al. 1989; Grupe et al. 2000;

Kohn et al. 1999; Krueger 1991; Lee-Thorp and van der Merwe 1987, 1991; Nelson et al.

1986; Nielsen-Marsh and Hedges 2000a; Schoeninger et al. 1989; Sillen 1989).

Traditionally, collagen has been preferentially used for the stable isotope investigation of ancient diet because it is less susceptible to diagenesis than is bioapatite 36

(see DeNiro 1985; Nelson et al. 1986). Because it is organic, however, collagen decomposes at a faster rate than bioapatite, which limits its use in palaeodietary studies to

Late Pleistocene and Holocene materials (Bocherens et al. 1991; DeNiro and Weiner

1988; Hare 1980; Tuross et al. 1988). For this reason, researchers began to investigate the potential of using the mineral portion of bone and tooth for dietary studies, but found that diagenesis could significantly alter the stable carbon isotope values of carbonate (Krueger

1991; Lee-Thorp and van der Merwe 1991; Lee-Thorp et al. 1989). For example, Hassen

(1975) argued that due to postmortem alteration, bone apatite could not be accepted as representative of the original biogenic carbon until an independent means of identifying altered bone was developed.

This position was challenged in the 1980s when Sullivan and Krueger (1981,

1983; Krueger and Sullivan 1984) concluded that properly pre-treated bioapatite yielded reliable 13C/12C ratios that consistently differed from those obtained from collagen.

Bioapatite has since been found to be a valuable tissue to sample for stable carbon isotope analysis (Coyston 1995; Lee-Thorp 2000), especially when coupled with the isotopic analysis of collagen (see Clementz et al. 2009). While bone bioapatite is prone to diagenetic processes as a result of its poorly crystalline structure, tooth enamel is nonporous and has a larger apatite crystal size and therefore is more resistant to postmortem alteration (Quade et al. 1992; Trautz 1967), although some contamination can occur (Schoeninger et al. 2003; Sponheimer and Lee-Thorp 1999). As such, tooth enamel bioapatite has also been sampled for dietary studies. A number of studies have noted that the utilization of effective pretreatment procedures for collagen (Brown et al.

1988; Chisholm et al. 1983; J0rkov et al. 2007; Stafford et al. 1988; Tuross et al. 1988) and bioapatite (Garvie-Lok et al. 2004; Koch et al. 1997; Krueger 1991; Lambert et al. 37

1989; Lee-Thorp and van der Merwe 1987; Metcalfe et al. 2009b; Nielson-Marsh and

Hedges 2000b) can limit the effects of diagenesis. The techniques employed in this study are discussed below.

3.5.1 Assessing the Integrity of Preservation of Collagen

Several techniques have been developed to assess the degree of preservation and contamination of the organic component of bone. These include analyzing the collagen yield, the carbon to nitrogen (C/N) ratio, as well as the carbon and nitrogen content (wt %

C and wt % N, respectively) in a bone sample (Ambrose 1990; Grupe et al. 2000;

Schoeninger et al. 1989). All three techniques were used to determine the degree of preservation, and C/N, wt. %C and wt. %N were used to check for contamination of the

Caledonia bone collagen samples.

In order to calculate collagen yield, bone samples were weighed prior to treatment. The of collagen extracted was determined by subtracting the weight of the vial from the combined weight of the vial and sample after treatment, using the equation for collagen weight in Figure 3.2. Collagen yield (%) was then calculated by dividing the collagen weight by the original weight of the bone sample and multiplying by 100% (see equation for collagen yield in Figure 3.2).

Ambrose (1990) found that collagen yields below 4.5% in bones from Africa were not preserved well enough for isotopic analysis. However, other researchers have found

Collagen Weight = (Weight of Collagen + Vial) - (Weight of Vial)

Collagen Yield = (Collagen weight / Bone Weight) x 100%

Figure 3.2: Equations used to calculate collagen yield. 38

that collagen yields as low as 1% can provide reliable results (van Klinken 1999; White et

al. 1993). If, however, collagen yield drops below 0.5%, contamination becomes difficult to remove (van Klinken 1999:689). If collagen yields from the Caledonia samples fell

below 1%, but other indicators of preservation (C/N, wt % C and wt % N) were within the range acceptable for isotopic analysis, the samples were deemed sufficient preserved

for isotopic analysis.

The ratio of carbon to nitrogen atoms in bone collagen (i.e., C/N ratio) is measured using an elemental analyzer. Modern bone samples have a carbon to nitrogen ratio of between 2.9 and 3.6 (DeNiro 1985); as such, samples whose ratios fall within this range are considered to be well preserved. Although Prowse (2001) has suggested that diagenetically altered bone can still exhibit acceptable C/N ratios, when used in

conjunction with other indicators of preservation (i.e., collagen yield, % C and % N), it is

deemed to be an adequate indicator of sample preservation (Ambrose 1990, 1993;

Schoeninger et al. 1989; van Klinken 1999).

Modern animal collagen has nitrogen concentrations of 5.5% to 17.3% and carbon concentrations of 15.3% to 47% (Ambrose 1990:441). Collagen is considered to be well preserved if the carbon content (weight percent, wt % C) is above 13%, and the nitrogen content (wt % N) is above 4.8% (Ambrose 1990; Ambrose and Norr 1993). Samples with wt % C and wt % N concentrations below these values can still exhibit C/N ratios and collagen yields that farl within the range of well-preserved collagen as a result of the production of salts or other inorganic residues', especially if acids were incompletely removed after demineralization (Ambrose 1990). Thus it is suggested that wt % C and wt

% N provide the most unambiguous indicator of the degree of preservation of collagen (Ambrose 1990; Schoeninger et al. 1989). It is still best, however, to utilize multiple indicators to assess the preservation of samples.

Certain contaminants can also be removed using accepted pretreatment protocols.

Chisholm and colleagues (1983) proposed a method to improve the collagen extraction technique pioneered by Longin (1971), in which collagen is treated with sodium hydroxide (NaOH) after initial treatment with hydrochloric acid. The application of

NaOH removes more soluble contaminants from the environment, such as humic and fulvic acids, which can otherwise skew the results of isotopic analysis.

3.5.2 Assessing the Integrity of Preservation of Bioapatite

Methods have also been proposed to assess the degree of contamination of bioapatite in both bone and tooth samples (see Lambert et al. 1985, Rink and Schwarcz

1995), including the analysis of histological structures (Metcalfe et al. 2009b; Pfeiffer et al. 2000; Schoeninger et al. 1989), differing x-ray techniques (El-Kammar et al. 1989;

Lee-Thorp and Sponheimer 2003; Metcalfe et al. 2009b; Schoeninger et al. 1989), CO2 yield (Koch et al. 1997; Metcalfe et al. 2009b; White et al. 2007; Williams et al. 2005,

2009; Wright and Schwarcz 1996, 1998; Wright et al. 2010), and infrared spectroscopy .

(Nielsen-Marsh and Hedges 2000a; Sillen 1989; Sillen and Sealy 1995; see Chapter

Five). Additionally, the use of proper pretreatment procedures in order to remove contaminants (see Ambrose 1993; Garvie-Lok et al. 2004; Koch et al. 1997; Krueger

1991; Lambert et al. 1989; Lee Thorp and van der Merwe 1987, 1991; Nielsen-Marsh and

Hedges 2000b; Sullivan and Krueger 1981; but see Zazzo et al. 2004) can mitigate the effects of diagenesis on bone and tooth bioapatite samples used for dietary reconstruction. 40

Fourier transform infrared (FTIR) spectroscopy measures the absorption of infrared radiation by the sample at the vibrational frequencies of its component molecular bonds. In this way, infrared spectroscopy can be used to characterize the structural sites of a compound. Changes in the crystalline structure of powdered bone or tooth enamel samples due to diagenesis are evaluated by measuring their absorbance spectra. The crystallinity index (CI) reflects both the relative sizes of the crystals in apatite as well as the degree to which the atoms are ordered (Weiner and Bar Yosef 1990), which indicate the degree of alteration. The application of FTIR to determine the preservation of bone and tooth bioapatite is now commonly implemented by researchers (see, for example,

Lee-Thorp and Sponheimer 2003; Nielsen-Marsh and Hedges 2000a; Wright and

Schwarcz 1996).

To assess the degree of diagenetic alteration of bioapatite in bone and tooth, samples are prepared following standard methods (Wright and Schwarcz 1996; see

Chapter Five). The CI for each sample was calculated to assess the degree of recrystallization using the method outlined by Weiner and Bar Yosef (1990). This is done by drawing a line on the absorbance spectra between the lowest "local" points on the chart to correct the baseline, which represents the background for the measurement. Then the distance between the baseline and points A, B, and C are measured, and the CI is calculated using the formula presented in Figure 5.1. According to Shemesh (1990), low

CI is between 3.0 and 3.7 and high CI is above 4.0. If the crystallinity indices are below the range of 2.8 to 4.0 found for modern human bone (Wright and Schwarcz 1996), the bone has not been significantly recrystallized. Similarly, if the tooth enamel samples have

CI values within the range found for modern tooth enamel samples (3.8 and 4.4;

Keenleyside et al. 2011), they are considered to be well preserved. 41

0.9

0.7

oa> s •e |0,

0.3

850 750 650 550 450 Wavenumber (cm"1)

Figure 3.3: Typical FTIR graph and method for calculating crystallinity index (CI) (after Shemesh 1990:2434; Weiner and Bar-Yosef 1990:191).

According to previous studies (Prowse et al. 2007; Weiner and Bar-Yosef 1990;

Wright and Schwarcz 1996), it is expected that archaeological bone will exhibit greater

CI than modem bones due to "diagenetic processes involving the growth of larger crystals at the expense of smaller ones" (Weiner and Bar-Yosef 1990:191). In addition, tooth enamel is expected to be more resilient to diagenetic alterations due to the higher rigidity of its crystalline structure when compared to bone (Koch et al. 1997; Lee-Thorp and van der Merwe 1987, 1991). The FTIR results for bone and tooth bioapatite samples from the

Caledonia collection are discussed in Section 6.2.1.

3.6 Chapter Three Summary

To summarize, C3 and C4 plants utilize two isotopically distinct photosynthetic pathways which can be used to interpret the diet of their consumers using stable carbon isotope analysis of both collagen and bioapatite samples. In areas where overlapping

CAM plants or marine resources are consumed, stable carbon isotope ratios alone cannot distinguish between multiple resources. In such cases, stable nitrogen isotope ratios in 42 collagen can indicate consumer trophic level, and the collagen-to-bioapatite spacing of carbon isotope ratios can indicate the degree of consumer carnivory.

Bone consists of an organic protein structure and an inorganic phase known as bioapatite from which collagen and structural carbonate, respectively, can be extracted and subjected to isotopic analysis for dietary reconstruction. This is because the tissues of a consumer reflect the isotopic composition of their diet (DeNiro and Epstein 1978, 1981;

Hare 1980; Macko et al. 1982). If protein consumption is adequate, the stable isotopes of carbon and nitrogen in collagen will be preferentially routed from protein in the diet, whereas the stable carbon isotopes in bioapatite will originate from the whole diet

(Ambrose and Norr 1993; Krueger and Sullivan 1984; Tieszen and Fagre 1993).

There also exist multiple factors that can complicate the interpretation of diet using stable carbon and nitrogen isotope analysis (i.e., diagenesis, trophic level effect in

815N values, or the canopy effect on S13C values). However, a thorough understanding of how and why these influence delta values can minimize the impact they will have on dietary interpretation. Furthermore, proper detection and pretreatment measures can mitigate the impact of diagenesis on isotopic interpretations of diet. 43

CHAPTER FOUR

ANCIENT MAYA FOOD RESOURCES

4.1 Indirect Evidence of Ancient Maya Food Consumption

Researchers employ multiple lines of evidence in order to interpret the use of food resources at archaeological sites. Such evidence includes excavated plant and animal remains, artefacts associated with food production, ethnohistoric and ethnographic accounts and linguistic studies. In addition, architectural evidence of agriculture, depictions of plants and animals in codices and on ceramics, as well as pathology related to health and nutrition in human skeletal remains are examined in order to suggest what resources may have been used (see Schoeninger and Moore 1992). It is necessary to review these sources and understand what food resources would have been available (i.e., the menu) before the actual foods consumed (i.e., the meal) can be assessed using stable carbon and nitrogen isotope analysis of the human skeletal remains from Caledonia.

Below, each line of evidence mentioned above will be discussed in terms of food resources from different sites throughout the ancient Maya region. Specific evidence of diet from Caledonia will then be discussed, followed by a review of stable carbon and nitrogen isotope studies that have been used to infer ancient Maya diet.

4.1.1 Floral Resources

The preserved remains of plants, both microscopic (i.e., phytoliths or pollen) and macroscopic (most commonly seeds), indicate the types of plants that would have been available in a given region. Interpreting the contribution of plant remains to ancient Maya diet is, however, limited by two primary factors: (1) preservation; and (2) archaeological sampling and recovery. Many parts of maize (Zea mays), preserve well, and thus it is not 44

surprising that this plant has been recovered from many Maya sites where flotation and/or

pollen samples have been collected and analyzed (Crane 1996; Lentz 1991; Miksicek

1991; Turner and Miksicek 1984). Conversely, crops such as beans, squash and manioc

do not preserve well, and thus are difficult to identify in the archaeological record (see

Carbone and Keel 1985; Pohl et al. 1996). The preservation of plant remains is further

hampered by cultural practices such as cooking methods which remove evidence of plant

remains from sites. For example, beans are usually prepared by soaking followed by

boiling and consumption of the entire bean and, thus, they are not likely to be well

represented in the archaeological record (see Crane 1996:267). It has been only within the

last few decades that the recovery of archaeobotanical remains from sites has been

incorporated into archaeological projects in the Maya region. Excavations at ,

Belize from 1977 to 1979 were the first to collect archaeobotanical remains

systematically (Crane 1996). Subsequent studies have found that a combination of

flotation and dry screening techniques maximizes the quantity and variety of plant

remains that can be recovered from a site (Hageman and Goldstein 2009; Lentz 1999;

Turner and Miksicek 1984). Appendix B lists a number of botanical remains recovered

from ancient Maya sites. Although this list is not exhaustive, it serves to demonstrate the

variety of resources utilized by the ancient Maya. For a more detailed discussion of the

types of plants the ancient Maya made use of, the reader is directed elsewhere (Atran and

UcanEk' 1999; Lentz 1999).

Artefacts, architecture and linguistic studies can provide evidence of the variety of plant remains that may have been consumed by the ancient Maya as well. For example,

manos and metates (equivalent to mortars and pestles) are commonly associated with the

processing of plant remains, primarily maize (see Pohl et al. 1996:365), but also other 45 plants such as cotton seeds (Lentz et al. 1996) at Maya sites. In addition, analysis of ceramic serving vessels has been used to indicate the types of food they contained (see

LeCount 2001; Lentz et al. 1996), and reveal how they were used in the preparation of foods on a daily basis (see Cheetham 2010; Powis 2004a). Similarly, linguistic data has been used to estimate the approximate date at which beans were introduced into the diet

(roughly 1400 B.C; Brown 2010).

Studies of modem Maya people living in Central America can also provide analogies with which to hypothesize the activities of their ancestors in the past. For example, researchers originally assumed that the ancient Maya employed slash-and-burn, or swidden, agriculture to grow crops for consumption because this type of farming predominates in the Southern Lowlands today (Morley 1946:141; see also McGee 2002).

This system of agriculture involves clearing an expanse of forest, cultivating the land for several years until the soil is depleted, and then leaving the land fallow while clearing another expanse of forest for cultivation. However, upon reviewing population estimates from large sites in the Maya Lowlands, it was determined that swidden agriculture would not have been sufficient to support the population (Hammond 1978; Turner 1978).

Researchers have since identified several agricultural systems that were used by the ancient Maya, including terraced (i.e., Healy et al. 1980, 1983) and raised fields (i.e.,

Siemens and Puleston 1972), which would have increased the production of food to that required by a large population. In the Maya Lowlands, the constmction of terraces is most likely to be"associated with the Late Classic period, when population levels in the area were at their highest (Healy 1990; Healy et al. 1983; White et al. 1993). The ancient

Maya were also likely relying on household gardens for subsistence (Hammond 1978).

For example, architecture preserved by a volcanic emption at Joya de Ceren, El Salvador, 46 has revealed kitchen gardens associated with houses where maize, beans, squash, seed and root crops would have been grown (Sheets 2000), and plaster casts revealed the

locations of garden fences and food storerooms (Lentz et al. 1996).

Maize was an important plant to the ancient Maya, both culturally and spiritually, but also in terms of subsistence (Hellmuth 1977; Marcus 1982; McGee 2002; Tozzer

1941). This plant remains important among modem Maya descendents, although it has become less prevalent in the diet because of increased access to a variety of foods resulting from modem global exchange in foodstuffs (van der Merwe et al. 2000).

Because of the poor preservation of plant material discussed above, maize tends to be the most frequently preserved plant, and thus was thought to be the only staple plant food consumed by the Maya. Upon review of ethnohistoric and ethnographic accounts, however, it has been found that the Maya utilized a variety of plant species, both domesticated and wild (Lundell 1933, 1938; Nations and Nigh 1980; Steggerda 1941;

Tozzer 1941). Some ethnographic studies have focused on medicinal plants (Ankli et al.

1999) that are not considered food items, but nonetheless were consumed, and thus would influence the isotopic composition of human skeletal material. It is likely that the majority of medicinal plants follow the C3 photosynthetic pathway, which would be reflected in the carbon isotope signature within the bones of individuals (see Section 3.3.1).

Additionally, the use of food in rituals practiced by the modem Maya (see Stress 2010) can help us infer via analogy how their ancestors may have similarly used food sources

(Christenson 2010). Finally, an understanding of food preparation techniques amongthe modem Maya, such as the preparation of the jute snail for consumption (Healy et al.

1990), can reveal how resources may have been used in the past.

According to Hawkes (1998), 49 plant species were domesticated in 47

Mesoamerica, including maize, beans, squash, spices, fruits and fibre plants.

Ethnohistoric and ethnographic studies indicate that the traditional Maya diet consisted of maize, beans, squash, chillies and cacao (Farriss 1984; Lentz 1999; Redfield and Villa

Rojas 1962; Villa Rojas 1945). Furthermore, several varieties of maize were cultivated by the ancient Maya (Tuxill et al. 2010). Researchers have found that maize and manioc were introduced into the Maya Lowlands before 3000 B.C., possibly as early as 3400

B.C., although maize cultivation became common only after 2400 B.C. (Pohl 1990b; Pohl et al. 1996). Chili pepper and cotton were introduced by roughly 1700 B.C. (Hammond

2001), followed by squash and bottle gourd around 1500 B.C. (Pohl et al. 1996).

Research has revealed that the consumption of certain plants varied over time. At

Cerros, for example, the frequency of maize and squash remained constant over time, but the consumption of the tree fruits nance and coyol palm was elevated during the

Preclassic period (Crane 1996). Furthermore, the use of particular species of plants depended on the environment and the ability to obtain them. For example, palms grow primarily on the coast; therefore they would be more common at coastal sites, and if found inland, they were most likely traded (Lentz 1991). Thus, the types of plants consumed by the Maya at Caledonia will likely reflect both the local ecology as well as the degree to which Caledonia traded with other centres over time.

4.1.2 Faunal Resources

In addition to floral remains, faunal remains recovered from archaeological sites can provide information on the types of protein sources consumed in the past. A variety of faunal remains, including deer, peccary, and dog, as well as birds, freshwater and marine molluscs and fishes, and reptiles such as turtle and crocodile, have been recovered 48 from ancient Maya sites. A list of animal species from a selection of ancient Maya archaeological sites can be found in Appendix B. Detailed descriptions of different animal resources utilized by the ancient Maya can be found elsewhere (Carr 1986, 1996;

Crane and Carr 1994; Emery 2004; Hamblin and Rea 1985; Powis 2004b).

Depictions of animals in ceramics, murals and other artwork, can provide insight into the types of species deemed important in Maya culture. It is important to recognize, however, that some of these images may reflect deities or other culturally significant figures rather than species related to subsistence. For example, depictions of deer have been identified in one of the few surviving Maya books, The Madrid Codex, as well as on pottery (see Carr 1996), and have been interpreted as either ritual (Pohl 1981:516) or possibly hunting scenes (Carr 1996; Franco 1969).

Like plant remains, the remains of animals at archaeological sites in specific contexts can provide information regarding diet. For example, the most frequent species in assemblages of animal remains from many ancient Maya sites was the white-tailed deer (Carr 1986; Olsen 1972, 1978; Pohl 1990a; Pollock and Ray 1957; Scott 1980;

Willey et al. 1965:523; Wing 1975; Wing and Scudder 1991; Wing and Steadman 1980), which suggests that it was a desired food resource. Conversely, there was a low dependence on deer on Island; instead, bird bone is predominantly recovered from elite contexts, indicating a connection between social status and avian species

(Hamblin and Rea 1985). This suggests that, as with plant resources, the type of faunal resources exploited depended on the local ecology of the area surrounding the site.

Dogs were domesticated in North America roughly 4000 B.C., and the turkey was domesticated about 2000 B.C. (Valadez 1999). The remains of both of these animals have been recovered from archaeological sites and ethnohistoric accounts suggest that they 49 were consumed on a regular basis (Diaz del Castillo 1927; Tozzer 1941; Ximenex 1967).

These accounts also suggest that deer were semi-domesticated by the ancient Maya, and this has been investigated by several researchers using faunal remains (Carr 1996;

Masson and Lope 2008; Pohl 1983). Isotopic analyses of the bones of peccary, deer and dog have found that all three species consumed maize at several sites, although the extent varied by site and time period, and the degree of human involvement is still unclear

(Emery et al. 2000; van der Merwe et al. 2000; White and Schwarcz 1989; White et al.

1993, 2001a).

Historically among the Maya the consumption of meat was a sign of elevated social status and "was a way of asserting class distinctions" (Pohl 1983:147; see also

Emery 2003; Pohl 1985). Therefore, it is possible that the use of meat to distinguish class also existed during the Classic and even into the Preclassic periods. It is likely, however, that the consumption of animal resources changed over time. At Cerros, for example, the consumption offish species, with the exception of parrot fish, decreased over time, and an increase in the consumption of turtle, white-tailed deer and peccary was observed during the Preclassic period (Carr 1986). This variation in the utilization of animal resources over time is found at other sites as well. For example, ocellated turkey

(Meleagris ocellata) is more abundant at the Preclassic to Postclassic Northern Lowland

Yucatecan sites (Hamblin 1984; Pollock and Ray 1957; Wing 1981), than at Southern

Lowland sites such as Seibal, Altar de Sacrificios, Cerros, and (Carr 1985; Pohl

1976; Wing 1975). In addition, Late Preclassic sites on the coast, such as Cerros (Carr

1985), and Postclassic Cozumel (Hamblin 1985), were heavily dependent on reef and marine resources. Stable carbon and nitrogen isotope analysis can help reveal dietary 50 patterns at specific sites, and further situate a site such as Caledonia within broader spatial and temporal contexts within the Maya Lowlands.

4.1.3 Skeletal and Dental Pathology

It is important to document evidence of disease in the skeleton and dentition of an individual when investigating their diet. The presence of pathological lesions related to poor nutrition can be used as indicators of diet because of the synergistic relationship between nutrition and disease (Armelagos 1994; Cohen and Armelagos 1984; Larsen

1987; Martin et al. 1985; Ortner and Aufderheide 1991). Several conditions that have been linked to subsistence in the Maya region, including porotic hyperostosis, cribra orbitalia, linear enamel hypoplasia, dental caries, dental calculus and tooth wear, are discussed below.

Porotic hyperostosis is characterized by the expansion of the diploe between the inner and outer tables of the skull, followed by a thinning of the outer table that results in porous lesions most commonly present on the parietal and occipital bones. Although in the past similar lesions on the roof of the orbits have been grouped together with lesions on the cranial vault (see Stuart-Macadam 1992), here the former will be referred to as cribra orbitalia, and the latter will be referred to as porotic hyperostosis. Both conditions cannot occur in adults because hemopoietic marrow is gradually replaced by fatty marrow with age, although the healed childhood lesions can persist into adulthood (Stuart-

Macadam 1985). The exact etiologies of porotic hyperostosis and cribra orbitalia are not well understood, although they have been related to iron deficiency anemia resulting from poor nutrition (Huss-Ashmore et al. 1982; Larsen 1987; Ortner 2003:359-382; Palkovich

1987; Stuart-Macadam 1985, 1988, 1992; Wright and Chew 1998). However, they may 51 alternatively result from anemia caused by parasitic infection (Holland and O'Brien

1997), or megaloblastic anemia (Walker et al. 2009). Because there is a lack of evidence for genetic haemolytic anemia in the New World, porotic hyperostosis has been attributed to iron deficiency anemia among the Maya (El-Najjar 1977; Saul 1977; Stuart-Macadam

1992). Furthermore, this condition may be influenced by a number of factors such as individual physiology (i.e., sex or age), and may also be caused by other disease processes, for example metabolic disorders such as vitamin C deficiency (Stuart-

Macadam 1989; Stuart-Macadam and Kent 1992). Similarly, a number of causes have been proposed for cribra orbitalia including inflammation and osteoporosis (Wapler et al.

2004).

Both porotic hyperostosis and cribra orbitalia have been identified in skeletal collections from many Maya sites, and in many cases the lesions have been attributed to dietary anemia resulting from a maize-based diet (D. Chase 1997; Saul 1977; Scherer et al. 2007; Whittington 2003; Whittington and Reed 1997; Wright and Chew 1998; but see

Williams et al. 2009; Wright and White 1996). The complicated etiology of porotic hyperostosis is, however, gaining recognition (see Wright and White 1996 for a review), and some studies have suggested non-dietary explanations for the presence of porotic hyperostosis and cribra orbitalia in Maya skeletal samples, such as anemia resulting from infectious disease (White et al 1994; Wright and Chew 1998). Although the precise etiology of these conditions remains difficult to determine, the presence of porotic hyperostosis and cribra orbitalia among the Caledonia skeletons will be recorded for comparison with the isotopic results to investigate whether a link between diet and porotic hyperostosis can be established at this site. 52

Linear enamel hypoplasia represents a developmental arrest of enamel or underlying tissue during the process of crown formation and has been related to a variety of systemic disturbances, including malnutrition and the presence of infectious disease during childhood (Huss-Ashmore et al. 1982; Kreshover 1960; Larsen 1987; Rose et al.

1985). Linear enamel hypoplasia is relatively common in the Maya region, and has been identified at a number of sites such as Caracol (D. Chase 1994, 1997) in Belize, several sites in the Pasion region of Guatemala (Wright 1997b, 2006), and at the Highland site of

Iximche, Guatemala (Whittington 2003). Defects are more commonly found on the anterior teeth (incisors and canines), and severe periods of childhood stress would be necessary for defects to appear on the posterior teeth (Wright 1997b).

Conditions of the dentition, including dental calculus (mineralized plaque) and dental caries can also indirectly indicate diet. Because these conditions are interrelated, they will be discussed together here. Although not itself a disease process, dental calculus, or calcified plaque (deposits of bacteria on the tooth surface), has been used as an indicator of a diet high in protein (HiUson 1979). In the Maya region, Evans (1973) was the first to use dental calculus to indicate diet, although it was argued that carbohydrates were the causative factor for the formation of dental calculus in that study.

It should be noted, however, that several factors, such as general diet, cultural practices, habitual activities and taphonomic processes will have an effect on the formation and subsequent visibility of calculus on the dentition (Hillson 1979; Lieverse 1999; Magennis

1999). Thefefore, it is best to use the presence and degree of calculus in the dentition in conjunction with more reliable indicators of diet.

Higher frequencies of caries (age-dependent destruction of enamel, dentine and/or cementum as a result of acid produced by bacteria in dental plaque) have been related to 53 diets high in carbohydrates and refined sugars (Bibby 1961; Navia 1994; Sheiham 1983).

In ancient Maya skeletons, high rates of these lesions have been attributed to diets that depend largely on maize (Evans 1973; Huss-Ashmore et al. 1982; Larsen 1987; Ortner

2003; Powell 1985; Tayles et al. 2000; Whittington 1999). It appears that the consistency and frequency of carbohydrate consumption are more important in the expression of caries than the actual amount of carbohydrates consumed (Bibby 1961; Magennis 1999;

Whittington 1999). Furthermore, although at the individual level dental calculus and caries can occur on the same tooth, at the population level there may be an inverse relationship between caries and calculus frequencies (HiUson 1996:260; see also Evans

1973). This is logical considering that high protein diets tend to produce a higher incidence of calculus and a lower incidence of caries, while a diet high in carbohydrates will result in the opposite trend (HiUson 1979). Similarly, it is possible that high levels of protein and fat in the diet may provide some protection against caries (Bowen and

Pearson 1993; Costa 1980; HiUson 1996; Magennis 1999). The ancient Maya, however, relied primarily on plant foods (Coe 1994), and it is likely, therefore, that they consumed more plant carbohydrates and less animal protein and fat which could have protected against caries. Dental caries and calculus have been identified in several Maya samples

(Magennis 1999; White 1988; White and Schwarcz 1989; Whittington 1999), and they are expected to be present in the Caledonia collection as well. If so, they will be compared to the results of stable isotope analyses to assess dietary practices at the site.

The final aspect of the dentition to be discussed here is tooth wear, which is a result of use rather than a disease process. Dental wear has been proposed as a means for estimating the age of a skeleton, although this method is not particularly useful in past populations because of dietary influences (see Walker et al. 1991). In general, high levels 54

of tooth wear reported in agricultural populations have been attributed to the consumption

of coarse cereal grains (i.e., maize) and the use of grinding stones (i.e., manos and metates) which add grit to the diet that can subsequently cause significant wear to the teeth (Powell 1985). While it is assumed that maize was the primary plant to be ground using a mano and metate and thus high rates of tooth wear indicate a diet high in maize

(Glassman and Garber 1999; Whittington 2003), there is evidence that many plant species were processed in this way by the ancient Maya (see Lentz et al. 1996). The presence of

antemortem tooth wear is therefore more indicative of how plant foods were processed by the Maya, rather than the specific types of plants that were consumed.

Finally, although some researchers have argued that dental wear was slight among the ancient Maya because of the lack of abrasive foods in the diet due to the high level of maize processing (Scherer et al. 2007), it has been identified in varying degrees in other ancient Maya skeletal collections and is expected to be present in the Caledonia sample.

Because tooth wear can impact the visibility of dental caries in the dentition, which can provide information regarding diet (see above), it will be assessed in order to understand the impact of this condition on dietary interpretations at Caledonia.

4.2 Food Resources at Caledonia

4.2.1 Modern Subsistence Resources at Caledonia

Due to the close proximity of Caledonia to the Macal River and its location in an ecologically transitional zone, the environment around the site is slightly different from that of the greater Chiquibul region. There appears to be a divide in vegetation and geology in the region between the site and the river (Awe 1985:19). The region closest to the site is characterized by a large variety of floral species, including mahogany, sapodiUa 55 and ramon, whereas the vegetation of the hillside forest gives way to dense shrub with several giant fig trees (Ficus galbrata) located near the river bank (Awe 1985:19).

A full description of the flora and fauna observed by the excavation team can be found in Awe (1985). Aquatic faunal species included crocodile, iguana, catfish, gar, carp and minnow as well as fresh-water mussel, and jute river snails. The majority of these species would have been edible; however, Awe (1985:19) claims that the first three would have provided more food per unit due to their large size. Terrestrial species included a variety of birds such as scarlet macaw "and ocellated turkey, and mammals included tapir, peccary, savannah deer, puma, jaguar and howler monkeys. Many of these species would have been exploited by the ancient Maya for food as well as for their materials (i.e., deer bone awls; jaguar, puma, tapir, and possibly peccary skins; Awe 1985:19).

4.2.2 Archaeological Evidence of Dietary Resources at Caledonia

Mano and metate fragments were excavated from a variety of contexts at

Caledonia (Awe 1985) and were likely used in the processing of plant foods. A ceramic vessel depicting a deer was found in Burial #1 in Str. A-1 and dated from the Early to .

Late Classic (Awe 1985:138-140; Figure 4.1). Finally, several deer bone awls were excavated from different contexts at the site (Awe 1985:359-360) and, although these are not directly related to diet, it is assumed that deer were consumed prior to the conversion of their bone into tools.

No botanical investigations were undertaken at Caledonia and, therefore,- studies from sites in the surrounding area are discussed here. Virtually no pollen was recovered from terraces at Caracol, or those located within the Cayo district (Healy et al. 1983:407).

Large quantities of minute charcoal fragments in the lower levels of terraces at Mountain 56

Figure 4.1: Deer depicted on Vessel 8 recovered from Burial #1 in Str. A-1 from Caledonia, dating to the Early to Late Classic transition (A.D. 450 to 650) (after Awe 1985:138-140; Healy et al. 1998:267).

Cow, Zayden Creek and Caracol have been interpreted as evidence that the terrace plots had an agricultural function (Healy et al. 1983:407). Preserved pollen, including maize

(Zea mays), ragweed (Ambrosia) and other Compositae (weeds that are commonly associated with the disturbance of natural vegetation and cultivated fields) were recovered from a core taken from the large Southern reservoir at Caracol (Healy et al. 1983:407).

Furthermore, isotopic analysis of the soils from terraces near Caracol exhibited enriched

813C values, indicating the cultivation of maize in these areas (Webb et al. 2004; see also

Webb et al. 2007).

Faunal remains were recovered from several areas of the site, some of which may have been due to subsistence activities, and others may have been related to ceremonial use. Worked marine shell recovered from the site formed pendants and may have been incorporated into a mural at Caledonia (Awe 1985:362-371). Among the human remains recovered from Caledonia were several non-human bone and shell fragments (see

Appendix A). For example, among the remains from Burial #1, a long bone fragment resembling that of an ungulate (possibly a deer) was identified, although a more specific 57 identification was not possible due to poor preservation. An antler fragment from an unknown species of deer and the claws of a crayfish were excavated from several contexts at the site (Awe 1985:78). Unmodified shell from freshwater mussels

(Nephronaias ortomanni) and snails (Pomacea and Pachyus) were present in all time periods and were likely obtained locally. Although the mussel shells may only represent ornamentation, it is possible that the meat was consumed and the shells perforated for use as ornaments (Powis 2004b) and, it is likely that the jute were consumed as well (Healy et al. 1990). All marine shell recovered from Caledonia dated from the Early or Middle

Classic Periods and would have been imported to the site (Awe 1985:377), likely for ornamentation. Marine shell from the Caribbean Sea and Atlantic Ocean included bivalves and gastropods, and bivalves and conchs were also obtained from the Pacific

Ocean, although no marine fish remains were recovered from Caledonia (Awe 1985).

4.3 Stable Carbon and Nitrogen Isotope Studies in the Maya Region

4.3.1 Isotopic Food Webs in Mesoamerica

In order to interpret diet using stable carbon and nitrogen isotope analysis of human tissue, the results must be compared to the isotopic composition of various resources available to the population under investigation. Site-specific models of isotopic variability in foods (i.e., food webs) are created by analysing the stable carbon and nitrogen isotope signatures of modem plants and animal resources from within the site vicinity (correcting for carbon output from the industrial revolution, see Section 3.3.6), and from archaeological animal remains preserved at the site itself. These values are used to create a food web, with which the human values are compared in order to interpret which resources they consumed in the past. 58

It is important to create food webs specific to a particular area because of the range of variation in stable carbon and nitrogen isotopes worldwide; a food web created for one area may not reflect the stable isotope values of resources in another. Similarly, slight variation can occur within a single region. For example, Figure 4.2 depicts two food webs from the Southern Maya Lowlands; the diagram on the right was developed from archaeological and modem resources at sites in Belize (Coyston et al. 1999:225), and the diagram on the left represents archaeological and modem resources from sites in the Pasion Valley of Guatemala (Wright 1997a: 187). This illustrates'the geographical variation in resource isotope values within a single region, in this case the Southern Maya

Lowlands, but is also dependent on what resources and tissues were sampled by the researcher (see Coyston et al. 1999:225; Wright 1997a: 187).

No site-specific stable carbon or nitrogen isotope values were obtained for resources available to individuals who resided at Caledonia. Instead, a composite food web was created using stable carbon and nitrogen isotope data on resources that would have been available to the inhabitants of this site (see Figure 4.3). This food web is based on data from a number of sources (Keegan and DeNiro 1988; Norr 1991; White and Schwarcz

1989; White et al. 1993; Williams et al. 2009; Wright 2006) and the 813C values of modem plants and animals were corrected by +1.5%o to compensate for the effects of fossil fuels on stable carbon isotope ratios (Friedli et al. 1986; Keeling et al. 1979;

Marino and McElroy 1991). All values used to create this foodweb can be found in

Appendix C. Stable nitrogen isotope analysis could not be performed on all samples, and in several instances where the 815N values were more positive than expected (i.e., outliers) they were removed from the food web reconstmction (after Wright 2006). Most of the 8 Ccoi and 8 N values came from archaeological and modem samples at and 59

20 Lamanaiasti PadonVefle)* IR­ teiapeiaie mariae IS

14- fisbraeat freshwater fish J "mm/, 1 * tefreshaal aaimafe. !? 10 herbivore meat c* ! C4 tropical .reef fish 6 - saaSmeai =*? 4 - d=i write reef 2 - CAM& C4$asis sheliisk -j 1 1 i_ _i 1_ 1 1 1 L. -35 -30 -25 -20 -15 -10 -35 -3Q -25 -20 -15 -10 -5

Figure 4.2: Food webs created for two different areas occupied by the ancient Maya. Sites from inland Belize (Coyston et al. 1999:225) are illustrated on the left, and sites from the inland Pasion Valley region of Guatemala (Wright 1997a: 187) are illustrated on the right.

around ancient Maya sites, excluding the values from Keegan and DeNiro (1988; see

Appendix C), which were obtained from environments around the Bahaman Archipelago,

and some values obtained by Norr (1991) which came from various sites in Lower

Central America.

The values obtained from areas some distance from Caledonia should adequately

represent the resources that would have been available at the site. For example, although

1 "\

there is a fluctuation of marine 8 C values globally because the carbon isotopic

composition of HCOJ varies with pH and temperature (Faure 1986; Hoefs 2004), because

both Belize and the Bahamas are located on the Caribbean Sea, they share reef and

marine resources withsimilar isotopic signatures. Similarly, in both locations blue-green

algae with a very low 815N signature (~0%o) contribute a significant amount of nitrogen to

trophic reef and mangrove ecosystems (see Williams 2000:35). Therefore, the isotopic

values obtained for reef and lagoon resources off the coast of the Bahamas should

accurately represent the values in resources that would have been available to the ancient 18 i

16 - Inshore Fish 14 -

Freshwater Fish 12 - Land Crab Offshore Fish 10 - Reef/ < 8 - Terrestrial & risiuaime ^ Freshwater Fish Z 6 - 1 - 1 Animals Marine Decapods i VJ C? Plants Marine 4 - Invertebrates 1 2 - - Maize - \ Legumes Sea Grass and Algae 0

-2 -

/I _ —r t t i -35 • -30 -25 -20 -15 -10 -5 0

5"Ccol(%, ,VPDB) ure 4.3: Composite food web of inland and coastal resources that would have been available to the ancient Maya at Caledonia (data from Keegan and DeNiro 1988; Norr 1991; White and Schwarcz 1989; White et al. 1993; Williams 2000; Williams et al. 2009; Wright 2006). 61

Maya at Caledonia through trade with sites on the coast of Belize.

Plants were divided into several groups, including C3 plants, legumes, sea grass and algae, and maize (after Williams 2000; Williams et al. 2009) and specific species can be found in Appendix C. Terrestrial animals, including mammals (i.e., deer, tapir, peccary, agouti and dog) and iguana, were grouped with freshwater reptiles (i.e., turtle and crocodile) because they had similar stable carbon and nitrogen isotope values. All of the collared peccary, dog and one white-tailed deer had more positive 8 Ccoi values, likely the result of feeding on C4 foods (see Emery et al. 2000; van der Merwe et al.

2000; White et al. 2001a, 2004), but were still grouped with the other terrestrial animals to illustrate the overlap of animals who feed on both C3 and C4 plants.

The values of several land crab specimens were included because the remains of several unknown species of non-marine crab were recovered from differing contexts at

Caledonia (Awe 1985). The freshwater fish category included several species (see

Appendix C), although no fish remains were recovered from Caledonia (Awe 1985), perhaps due to poor preservation. Isotope data for the shells of jute snails and freshwater

13 15 mussels were available, but the S Ccoi and 8 N values of this material were much more negative than values obtained for jute snail flesh. As it is probable that only the flesh of these species was consumed, only the stable carbon and nitrogen isotope data of the meat of the jute were included in the food web (see Appendix C). Finally, marine animals were represented by a variety of species which were divided into the categories of reef/estuarine fish, inshore fish, offshore fish, marine invertebrates (i.e., marine snails, sea cucumbers) and marine decapods (i.e., shrimp, prawn, crabs, lobsters) after Williams

(2000; Williams et al. 2009; see Appendix C). 62

4.3.2 Stable Isotope Analysis of Ancient Maya Diet Review

Overviews of the available isotopic data from ancient Maya skeletal and dental

material are provided elsewhere (Tykot 2002; White et al. 2006b). These are summarized

in Appendix D along with more recent studies, and the general trends that have emerged

from this research are discussed below.

1 "\

The first isotopic analysis of ancient Maya skeletal remains evaluated the 8 Ccoi

and 815N values in bone collagen from a Classic period sample from , Belize in terms of age, sex, time period and social status (White 1986, 1988; White and Schwarcz

1989). The majority of subsequent studies have focused on the Southern Maya Lowlands, particularly sites in Belize (A. Chase and D. Chase 2001; A. Chase et al. 2001; Coyston

1995; Coyston et al. 1999; Gerry 1993, 1997; Henderson 1998, 2003; Metcalfe et al.

2009a; Powis et al. 1999; Scherer et al, 2007; Tykot et al. 1996; van der Merwe et al.

2000; White 1997; White et al. 1993, 1994, 2001a, 2001b, 2006a; Williams 2000;

Williams and White 2006; Williams et al. 2005, 2009), although some studies have

investigated the isotopic composition of diet at sites in the Peten region (Gerry 1993,

1997; Scherer et al. 2007; Wright 1994, 1997a, 1997b, 2006; Wright and Schwarcz 1996)

and highlands of Guatemala (Whittington 2003; Whittington and Reed 1998; Wright and

Schwarcz 1996, 1998, 1999; Wright et al. 2010), as well as sites in Honduras (Gerry

1993, 1997; Gerry and Kmeger 1997; Reed 1994, 1998, 1999) and the Yucatan peninsula

of Mexico (Mansell et al. 2006).

Many isotope data for ancient Maya diet come from skeletal material dating to the

Classic period (Gerry 1993, 1997; Mansell et al. 2006; Norr 1991; Reed 1994, 1998;

Scherer et al. 2007; White et al. 1993; Wright 1994, 1997a, 1997b, 2006), particularly that dating to the Late and Terminal Classic periods. This is because skeletal remains 63

dating to these time periods have the potential to answer questions relating to the Classic

period collapse (see Wright and White 1996), and are generally better preserved and have

larger sample sizes than skeletal collections from earlier periods. Nonetheless, the

majority of studies have examined multiple time periods at a single site (Coyston et al.

1999; A. Chase and D. Chase 2001; A. Chase et al. 2001; White 1988; White and

Schwarcz 1989; White et al. 1993, 2001b; Wright 1994; Wright and Schwarcz 1998), and

many have also focused on either the Preclassic, Postclassic and/or Historic periods

(Powis et al. 1999; Tykot et al. 1996; White et al. 1994; Whittington 2003; Whittington

and Reed 1998; Williams and White 2006; Williams et al. 2009).

Coyston (1995) was the first to employ 8 C in bioapatite for the reconstruction of

ancient Maya diet. This study sampled the skeletal collection from Classic period

Lamanai and and compared the results with the values for collagen obtained by

White (1988, 1989; White and Schwarcz 1989; White et al. 1993) to illustrate the

usefulness of bioapatite in dietary reconstmction if diagenesis is properly controlled for

(Coyston 1995; Coyston et al. 1999). Isotopic studies of ancient Maya diet have since begun to routinely sample carbon in collagen and bone bioapatite, and have begun to

sample tooth enamel bioapatite as well (A. Chase et al. 2001; Coyston et al. 1999; Gerry

and Kmeger 1997; Henderson 1998, 2003; Mansell et al. 2006; Metcalfe et al. 2009a;

Tykot et al. 1996; White et al. 2001b; Williams 2000; Williams et al. 2009; Wright and

Schwarcz 1996, 1998, 1999; Wright et al. 2010).

Cultural groups often have restrictions that regulate the distribution of food to different members of their society that are based on numerous economic, political, and ideological factors (Gumerman 1997). Most often such restrictions correspond to social divisions by age and sex, as well as the restriction of rare foods to those of elevated socio- 64 economic status (Wing and Brown 1979:11-12). Although this was also the case among the ancient Maya, it must be recognized that a pan-Maya model explaining the relationships between dietary change/stability and climate, ecological, political or economic factors has not been forthcoming (Wright and White 1996). Furthermore, temporal and regional pattems in dietary data can obscure more specific patterns related to sex, age group and social class. Nonetheless, stable carbon and nitrogen isotope analysis has the potential to provide information about social trends that cannot be identified using other methods. In order to understand these trends, a variety of factors including age and sex must be assessed in conjunction with each other. For example, at

Copan, Reed (1999) found that with increasing age, adult females consumed less C4 plant foods, whereas males of all ages consumed a similar diet. This suggests that some aspect of the social system at Copan resulted in age- and sex-based differences in diet that are undetectable using other lines of evidence (Reed 1999). Therefore, it is necessary to investigate dietary patterns related to age, sex and social status within the context of specific regions and time periods.

4.3.2.1 Variability by Age

The majority of isotopic studies investigating age-related dietary patterns among the Maya have investigated the timing of weaning. Breastfeeding infants exhibit 815N values that are approximately 2%o enriched relative to their mothers (Fogel et al. 1989).

Evidence of weaning is signalled by the decline in the stable nitrogen isotope ratios of young children such that they approach those of adult females (i.e., their mothers). For example, Wright and Schwarcz (1996) assessed stable carbon and oxygen isotope ratios in teeth that form at different ages to investigate the timing and duration of weaning. 65

Similarly, Williams and colleagues (2005) analyzed multiple isotopic indicators (i.e.,

IT n if io

8 CCoi, 8 Cbio, 8 N and A Cbio-coi) to further understand the weaning process at two

ancient Maya sites on the coast of Belize.

The transition from breastfeeding to solid foods has been investigated at several additional ancient Maya sites (Reed 1994; White and Schwarcz 1989; White et al. 2001b;

Williams 2000; Williams et al. 2005; Wright and Schwarcz 1996, 1999). These data, combined with studies of enamel hypoplasia (see White et al. 1994; Williams et al.

2005), suggest that Maya infants were breastfed from birth to at least three to four years of age. This has been substantiated using elemental analysis of the ratio of strontium to calcium (Sr/Ca) in teeth (see Song 2004).

Some studies have investigated age-related dietary patterns through a comparison of isotope ratios in subadult and adult remains, and have found varying trends. For example, at Pacbitun, subadults were found to have consumed less C4 foods than adults

(White et al. 1993), whereas at Copan the opposite trend was observed (Reed 1994).

Metcalfe and colleagues (2009a) found no differences between adults and juveniles, except that older adults had significantly higher 815N values than juveniles. Finally, at

Lamanai no age-related dietary differences were observed when subadult values were compared to those of adults (White 1986, 1988; White and Schwarcz 1989; White et al.

1994).

4.3.2.2 Variability by Sex

In addition to age, an individual's sex may have played a role in the types of foods deemed socially acceptable to consume. The investigation of sex-specific diets among the ancient Maya using stable carbon and nitrogen isotope analysis has been reviewed by 66

White (2005), and variable results have been published. For example, it has been

demonstrated that males consumed more C4-based foods than females at sites such as

Pacbitun (White et al. 1993) and (White et al. 2001b) in Belize, Copan (Reed

1994, 1999) in Honduras, as well as at Altar de Sacrificios, Seibal and Dos Pilas (Wright

1997a) in Guatemala. Alternatively, at in highland Guatemala, it is likely that

males consumed less C4 plants than females (Whittington 2003:301). Studies have also reported sex differences in the amount of animal protein consumed. For example, at Altun

Ha (White et al. 2001b) and Copan (Reed 1994) elevated S15N values suggest that males

consumed more meat than females. Finally, no significant dietary differences between the

sexes were observed at several sites throughout the Maya region (Gerry 1993, 1997;

Mansell et al. 2006; Metcalfe et al. 2009a; Tykot et al. 1996; White 1986; White and

Schwarcz 1989; Williams et al. 2009 Wright 1997a).

4.3.2.3 Variability by Social Status

In stratified societies, including the ancient Maya, different members of society

are subject to restrictions that regulate the distribution of food that are based on a variety

of social factors (Gumerman 1997; Wing and Brown 1979:11-12). Before dietary differences based on social status can be reviewed, the methods used to ascribe status to ancient Maya individuals must be acknowledged (see Becker 1992; D. Chase and A.

Chase 1992; Gillespie 2001; Pendergast 1992). One approach to assigning status is to use variables surrounding the burial (e.g., burial type and number and type of grave goods).

Some of these variables, however, are not necessarily adequate indicators of social status.

For example, it is generally assumed that burials located in the site core, tomb burials and elaborate grave goods belonged to individuals of high status, but relying on only one of 67 these characteristics to infer social status can be misleading (A. Chase 1992; A. Chase and D. Chase 1996, 2001; A. Chase et al. 2001; Scherer et al. 2007). One way that researchers have dealt with this is to assess as many of these variables as possible before assigning status (i.e., Gerry 1993; Scherer et al. 2007; White et al. 1993; Wright 2006).

Although socially-related dietary differences have been identified among the precontact Maya (see below), it is likely that the relationship between diet and status solidified during the Postclassic, and may have been intensified by the Spanish ideals imposed on the Maya during the Colonial era (see Gerry 1997:66). Thus, dietary trends related to social status may be more evident during later time periods. It is difficult to compare isotope data from different sites because researchers have tended to use different criteria to ascribe social status. Therefore the discussion of diet and social status presented below should be considered tentative, as a critical evaluation of the comparability of these studies is beyond the scope of this thesis.

Isotopic evidence of dietary variability between elites and lower status individuals has been found at many ancient Maya sites (A. Chase and D. Chase 2001; A. Chase et al.

2001; Coyston 1995; Coyston et al. 1999; Powis et al. 1999; Reed 1999; Scherer et al.

2007; White 1986; White and Schwarcz 1989; White et al. 1993, 1994, 2001b;

Whittington and Reed 1997; Wright 1997a, 1997b). The expression of these differences in diet based on status, however, is site dependent. For example, at Altun Ha (White et al.

2001b) and Pacbitun (White et al. 1993) higher status individuals consumed more maize and possibly maize-fed animals than lower status individuals, although no difference in

S15N values was noted at either site. Conversely, at Copan in Honduras elevated social status was associated with a more varied diet rather than access to protein or maize (Reed 68

1994), while at Chau Hiix in Belize it appears that the consumption of reef and marine resources distinguished the ruler from other residents at the site (Metcalfe et al. 2009a).

The stable carbon isotope analysis of bioapatite from both Pacbitun and Lamanai demonstrated that the significant differences in collagen data between elites and non- elites (White and Schwarcz 1989; White et al. 1993) were not similarly significant in bioapatite. This suggests that the dietary differences were related to meat consumption rather than whole diet (Coyston 1995; Coyston et al. 1999). Because the 815N values did not vary significantly between individuals, it was argued that the difference lay in the amount of meat consumed (Coyston et al. 1999). Similarly, the analysis of carbon isotope ratios in bone collagen suggested that a high status male buried in an Early Classic tomb at Lamanai had access to reef and marine resources, and more varied plant resources not available to lower status individuals at the site (White 1988; White and Schwarcz 1989).

Re-evaluation of the collagen data using carbon isotope ratios in bone bioapatite, however, revealed that this individual likely consumed C3-fed animals and/or freshwater fish in greater amounts than others at the site (Coyston et al. 1999). In contrast, individuals of elevated social status at Pacbitun appear to have consumed more C4-fed animals than lower status individuals (Coyston et al. 1999).

A similar conclusion was reached by Gerry (1993, 1997) during his analysis of several sites throughout the Southern Lowlands. Here, maize was important in all diets, though the highest level of society appears to have relied more heavily on C4 resources, while the lowest strata of society were the least dependent on this source of food and exhibited the most varied dietary protein. No significant difference was found in 815N values or A Cbio-coi, which together suggest an ommvorous diet consisting primarily of plant rather than animal resources. Gerry (1997) argued that if dietary discrimination 69 existed between the Classic Maya social classes, it was most likely qualitative rather than quantitative. In other words, it may be that individuals of elevated social status had privileged access to certain kinds of meat, such as puma, jaguar or deer (Gerry 1997; c.f.

Pohl 1985:137, 1994).

4.3.2.4 Temporal Variability

The investigation of subsistence change over time in the Maya region has focused on three primary areas of research. The first involves the study of general dietary trends at specific sites in relation to societal and political changes over time (Coyston 1995;

Coyston et al. 1999; Mansell et al. 2006; van der Merwe et al. 2000; White and Schwarcz

1989; White et al. 1993, 2001b). The second and third involve more specific questions relating to the impact of European contact on diet and health (White et al. 1994, 2006), and dietary evidence relating to the Classic Maya collapse circa A. D. 900 (Reed 1994; van der Merwe et al. 2000; Scherer et al. 2007; White 1986, 1988; White and Schwarcz

1989; Whittington and Reed 1997; Wright 1994, 1997a; Wright and White 1996), respectively.

Although temporal differences in the consumption of maize have been identified at most sites, these differences are not uniformly patterned (Coyston et al. 1999;

Henderson 1998, 2003; Tykot 2002; Tykot et al. 1996; Whittington 2003; Williams et al.

2009; Wright 1997b). Some researchers have proposed a general trend among the ancient

Maya towards increased maize consumption over time (Tykot 2002; White et al. 1993; van der Merwe et al. 2000). Others have identified a decrease in the consumption of maize foods during the Classic period (A. Chase and D. Chase 2001; A. Chase et al.

2001; Gerry 1997; White 1986, 1988). Still others have found a steady increase in maize 70 consumption until the middle of the Late Classic when a dramatic decrease is evident followed by another increase in maize consumption during the Postclassic period (White and Schwarcz 1989). Finally, at Piedras Negras the consumption of maize was more variable during the Early Classic relative to later periods (Scherer et al. 2007).

Temporal changes in diet can also be combined with dietary patterns related to age, sex, and social status in order to more clearly understand the change in diet. For example, at Dos Pilas and Altar de Sacrificios during the Terminal Classic, dietary differences by social status (as inferred from burial clusters) were not detected, although it was found that males consumed more C4 plants than females. Wright (1994, 1997a,

1997b) has suggested that this represents a shift from social inequality in access to food during the Classic to a less-stratified system during the Terminal Classic constrained by sex-specific patterns.

4.3.2.5 Regional Variability in Diet

The area inhabited by the ancient Maya encompassed a variety of environmental settings from coastal areas to the Guatemala highlands, and from the wet tropical forests to the arid region of the northern Yucatan. As a result, the ancient Maya employed a variety of agricultural and subsistence procurement methods. However, the presence of a particular form of agriculture in one region, for example terracing, does not necessarily mean it was employed in another region (Dunning 1996). Likewise, stable carbon and nitrogen isotope studies of ancient Maya human remains have revealed several region- specific dietary patterns. For example, during the Classic period, people in Guatemala and

Honduras remained more dependent on maize than those in Belize (Gerry 1997; van der

Merwe et al. 2000:34). It also appears that people in the Peten were more reliant on 71 animal protein than those in Belize or the Copan Valley, and that the legumes and other

C3 plant resources were more significant contributors to diet at Copan than at other sites in the Maya region (Gerry 1997).

This identification of regional diversity in food consumption among the ancient

Maya (Gerry 1993; Gerry and Kmeger 1997; Tykot et al. 1996; White 1997; Wright and

White 1996) has challenged the concept of a pan-Mesoamerican diet. For example, maize was an important component of ancient Maya diet; however, varying environments and time periods have resulted in significant variation in the quantity of maize consumed.

Furthermore, the exploitation of local environments played a significant role in the isotopic reflection of diet. For example, the heterogeneous environment at Lamanai as a result of its close proximity to the New River lagoon system may explain why the

13 individuals at this site consumed less C4 resources as evidenced by lower 8 Ccoi values compared to Pacbitun and Copan (White et al. 1993). Similarly, and not surprisingly, coastal sites in Belize have a distinct marine reef dietary component because of their proximity to these resources (Williams 2000; Williams et al. 2009). The local environment would have dictated the types of exploitable resources, and kin groups would have provided for themselves based on the types of available resources (Gerry

1997:63).

4.4 Chapter Four Summary

Although initially thought to have been nearly completely dependent on maize, it is now recognized that the ancient Maya utilized a variety of both plant and animal resources. While maize was certainly an important component in the diet, other plant resources were utilized, including cultigens such as beans, squash and chillies, as well as wild food and medicinal plants. Animal resources varied from freshwater species of 72 reptiles, snails and shellfish, to marine and reef resources including fish and shellfish, and terrestrial mammals such as deer, dog and peccary. When situated within the dietary context established by archaeological, zooarchaeological, palaeoecological, ethnohistoric, ethnographic, linguistic, iconographic and pathological studies, the analysis of stable carbon and nitrogen isotopes can provide direct evidence of the types of food consumed by ancient peoples. Such studies of ancient Maya skeletal remains have revealed dietary patterns related to age, sex, social status and time period, as well as regional variation in diet. Because limited evidence of dietary resources was recovered from Caledonia, the analysis of stable carbon and nitrogen isotopes in human skeletal remains will provide a better understanding of subsistence at this site. 73

CHAPTER FIVE

ANALYTICAL TECHNIQUES

5.1 Methods Employed in Skeletal Analysis

This skeletal collection has been previously examined by Dr. Herman Helmuth

(1985:414-429), and the study has verified and expanded upon his original observations. Neither I nor Dr. Helmuth had the opportunity to assess the individual from

Burial #2; therefore, this burial will not be discussed further.

The remains from each burial were first sorted by the individual designations

(e.g., Individual A from Burial 3) outlined in the original report (Helmuth 1985:414-429), as well as using labels written on some of the bones. In cases where the remains were fragmentary and commingled to the point that elements could not be associated with a specific individual, the minimum number of individuals (MNI) was calculated based on the premise that an individual can only have one of each bone (i.e., one right femur).

Thus, the presence of two or more of the same element suggests that a minimum of two individuals were present.

In order to interpret the results of stable carbon and nitrogen isotopic analysis meaningfully, it is necessary to provide contextual data for the skeletal collection sampled. When the age, sex, and pathology of the sampled individuals have been established, the isotopic results can add to interpretations of an individual's nutritional status, and can identify general dietary patterns related to age and sex. Here, the techniques used to estimate age, determine sex, and assess pathology are briefly described. Because not all skeletal elements were available for all individuals, different techniques were applied when needed. 74

5.1.1 Age Estimation Techniques

Methods for determining age at death can be divided into those that rely on changes in the growing skeleton and dentition and those that rely on the deterioration of the skeleton. The former are applicable to subadults and usually produce a narrow age range of one to three years. The deterioration of the skeleton begins between 18 and 25 years of age and continues throughout adulthood. Methods used to determine the age of adult skeletons thus provide wider age ranges, and are less accurate and more difficult to apply than methods developed for subadult remains. In addition, because different methods require specific bones which may not have been available from each individual recovered from Caledonia, as many methods as possible were employed to age each individual.

5.1.1.1 Age Estimation Techniques for Subadult Remains

Human dentition forms and empts at predictable stages during childhood growth.

The stages of tooth formation for the crown, root and apex of the deciduous mandibular canines, the deciduous mandibular molars and the permanent mandibular molars have been recorded by Moorrees and colleagues (1963a, 1963b), and are easily adapted to other tooth classes. Similarly, the emption sequence for the human dentition has been provided by Schour and Massler (1941:1154). Because the formation and emption of dentition occurs at predictable intervals during childhood, age estimations for subadults are more accurate than those for adults. When the dentition has fully formed and empted, however, only a minimum age estimate (i.e., 18 to 21 years) can be provided.

The degree of epiphyseal fusion (i.e., the union of epiphyseal caps and their associated diaphyses) can also be used to estimate the age at death of a skeleton. This 75

reflects the fact that epiphyseal union occurs at predictable stages for different bones, and results in a visible fusion line which is obliterated over time. If the bone is unfused, the

individual is considered to be the oldest age in the range or younger, and if a clear fusion

line is visible the individual is considered to be the youngest age in the range or older.

This technique is most useful for estimating the age at death of individuals between 10

and 25 years because this is the period during which the majority of epiphyses fuse.

However, the timing of epiphyseal fusion can be affected by factors such as diet, disease and population variation (Brothwell 1972; Mays 1998; Saunders 2000; Ubelaker 1978),

so it is best to use this method in conjunction with the formation and emption of the teeth.

The single subadult identified in the Caledonia collection was not sampled for isotopic analysis, and the methods described above were instead used to determine the minimum age of adults within the collection.

5.1.1.2 Age Estimation Techniques for Adult Remains

The morphological changes of the pubic symphysis of the pelvis are considered to be the most reliable criteria for estimating the age of adults, followed by morphological changes of the auricular surface (Buikstra and Ubelaker 1994:21). As age increases, the rugged bony ridges and grooves of these features are eroded and replaced by granular to fine-grained dense bone. For several reasons, including larger standard error in age ranges and a division by sex, the Suchey-Brooks (Brooks and Suchey 1990; Suchey and Katz

1986) method is considered more accurate than the Todd (1921a, 1921b) method for assessing sex from the morphology of the pubic symphysis. Thus, the Suchey-Brooks method is employed in this study. The morphology of the auricular surface, when present, was also assessed following the methods outlined by Lovejoy and colleagues (1985; 76

Meindl and Lovejoy 1989).

The degree of wear on the dentition can also indicate the general age of an individual. The method outlined in Smith (1984) for scoring the surface wear of the incisors, canines and premolars was employed to estimate the age of individuals with these available teeth. Individuals were considered to be young adults if their scores ranged from 0-3, middle-aged adults if their scores ranged from 4-6, and older adults if their score ranged from 7-8. Similarly, the Scott (1979) method for scoring the surface wear of the molars was implemented to estimate age at death when these teeth were available. Individuals were considered to be young adults if their molar wear ranged from

0 to 12, middle-aged adults if their score ranged from 13 to 27, and older adults if their score ranged from 28 to 40. Using tooth wear to estimate age at death is not as accurate as the techniques described above because diet can have an effect on the degree of dental wear (Smith 1984). Therefore, this method was used in conjunction with other aging techniques in order to estimate the age of each individual. When it was the only aging method employed, the estimate was considered tentative, and the individuals were classed broadly as adults.

In order to ensure accurate age estimations, as many of the methods described above were applied to each individual as preservation allowed. Using this combined information, the individuals were divided into five age categories as follows: subadult

(under 20 years), younger adult (20 to 35 years), middle-aged adult (35 to 50 years), and older adult (50 years or older) (Buikstra and Ubelaker 1994:9), and a general adult (over

20 years of age) category when a specific age range could not be estimated. 77

5.1.2 Sex Determination Techniques

When determining the sex of skeletal remains, features of the pelvis are the most accurate, followed by those of the cranium. In cases where these elements are absent or poorly preserved, alternative methods such as the diameter of the femoral head can be employed.

The pelvis (os coxa) is the most reliable indicator of sex in the skeleton as a result of morphological differences between males and females related to reproduction. The only two individuals who exhibited os coxae preserved well enough for sex determination were C2-3-A and C1-5-A. For these individuals, the presence of a ventral arc, subpubic concavity and/or an ischiopubic ramus ridge (Phenice 1969) were used to determine sex.

The greater sciatic notch was also used to assess sex, as the angle of this feature is significantly greater in females than in males (Singh and Potturi 1978). Here, sex was determined via visual examination of the angle of the greater sciatic notch compared to diagrams provided in Buikstra and Ubelaker (1994:18).

Only three skulls and a fragment of a fourth skull from Caledonia were preserved well enough for sex determination. Five features - the nuchal crest, mastoid process, supra-orbital margin, supra-orbital ridge/glabella and mental eminence - were assessed to determine sex using the skull. These features were scored from 1 (female) to 5 (male)

(Acsadi and Nemeskeri 1970: Figure 16).

In some cases where the pelvis and cranium are absent, sex can be established using postcranial remains. This involves the measurement of complete long limb bone lengths or the diameter of the heads of long limb bones. In most cases, however, the long limb bones of the individuals from Caledonia were fragmented and thus their lengths could not be measured. Sexual dimorphism of the femoral head is based on the fact that 78 the male axial skeleton experiences more weight than that of the female, which is

dissipated by the head of the femur (Williams et al. 1989). This method of sex determination is problematic for the following reasons: (1) the femoral head must be

intact to take an accurate measurement; and (2) the standards are based on specific modem populations (Asala 2001; Dittrick and Suchey 1986; Stewart 1979:120; Steyn and

Iscan 1997), which, may not provide accurate results when applied to an ancient Maya

skeletal sample. Thus, femoral head diameter was only used to estimate probable sex in cases where the cranium and pelvis were unavailable for observation.

5.1.3 Paleopathology: Nutrition and Disease

It is important to document evidence of disease in the skeleton and dentition when examining diet (see Section 4.1.3). For this study, the bones were visually examined with the aid of a magnifying glass in a well-lit room to facilitate identification of pathological lesions. Descriptions and photographs of all lesions observed were recorded (unpublished document in the possession of the author), and lesions were compared to descriptions and photographs in several texts and articles before a diagnosis was made (e.g., Bams 2005;

Brickley and Ives 2008; Buikstra and Ubelaker 1994; Katzenberg and Saunders 2008;

Mann and Hunt 2005; Ortner 2003; Palkovich 1987; Pinhasi and Mays 2008; Powell

1985; Roberts and Manchester 2007; Rose et al. 1985; Stuart-Macadam 1985; Stuart-

Macadam and Kent 1992).

Porous lesions consistent with porotic hyperostosis and cribra orbitalia were scored on the cranial vault and orbital plates, respectively. Lesions were recorded as present or absent and active or healed, and the severity of the porosity was scored according to Stuart-Macadam (1985). The dentition of the individuals sampled for isotope 79 analysis was examined for the presence of dental calculus, and if present, the location

(i.e., buccal, labial, interproximal surfaces) of the condition on each tooth was recorded

(see Appendix E). The severity of dental calculus was scored following the system provided in Brothwell (1981), as described by Buikstra and Ubelaker (1994:56).

As with calculus, the dentition of the individuals from Caledonia who were

sampled for isotope analysis was examined for dental caries. Lesions were recorded as present if penetration of the enamel coupled with discolouration was present on any tooth

surface, and the severity of the condition was recorded following the'system designed by

Moore and Corbett (1971), as modified by Buikstra and Ubelaker (1994:55). Linear enamel hypoplasia was scored as present if a visible enamel defect was present on the labial surface of the anterior teeth of individuals sampled for isotope analysis following the procedure outlined in Buikstra and Ubelaker (1994:56-58). The type of defect (i.e., linear horizontal grooves, linear vertical grooves, etc.) and number of defects were also recorded for each tooth. Finally, dental wear was recorded as described in a previous section (5.1.1.2).

5.2 Preparation of Bone Samples for Stable Isotope Analysis

5.2.1 Collagen

Table 5.1 lists the bones sampled for each individual in this study. Because of poor preservation, the same bone from each individual could not be sampled in all cases.

It has been demonstrated, however, that the mean carbon and nitrogen isotopic composition of collagen extracted from different bones of the same individual is not significantly different (DeNiro and Schoeninger 1983). Therefore, sampling collagen from different skeletal elements from each individual in the Caledonia collection will not 80

Table 5.1: Tissues Sampled for Stable Isotope Analysis

Burial # Individual Sample Code Bone Sampled 1 Unknown 1 A1-1-LM1 Left Fifth Metatarsal1 1 Unknown 2 A1-1-LM2 Left Fifth Metatarsal Unknown 3 A1-1-LM3 Left Fifth Metatarsal1

A C2-3-A Left Tibia1 Right Maxillary First Molar1 B C2-3-B Left Tibian1 Cranial Fragments Left Tibia2 C2-3-C 3 Right Femur Left Mandibular Third Molar3 D C2-3-D Right Tibia Mandible 5" C2-4-A Left Mandibular Third Molar3 Mandible B C2-4-B Right Mandibular Third Molar3 Mandible2. C2-4-C Right Mandibular Third Molar3 Unknown 1 C2-4-F1 Right FibulaT1 Unknown 2 C2-4-F2 Right Fibula Unknown 3 C2-4-F3 Right Fibula1" Unknown 4 C2-4-F4 Right Fibula1 Unknown 5 C2-4-F5 Right Fibula Unknown 6 C2-4-F6 Right Fibula1" Unknown 7 C2-4-F7 Right Fibula1 Rib Fragments C1-5-A Left Tibia1 Mandibular Right Second Molar3 Unknown 1 U-RM1 Right Fifth Metatarsal1 Unprovenienced Unknown 2 Right Fifth Metatarsal1 Burial U-RM2 Unknown 3 U-RM3 Right Fifth Metatarsal1 Collagen and bone bioapatite sampled fromthi s bone, 2Only collagen sampled from this bone. 3Only bioapatite sampled fromthi s bone/tooth. 4Collagen and bone bioapatite samples removed from analysis because of problems during preparation; an alternative sample of each was prepared. 5Collagen and bioapatite samples removed from analysis because of problems during preparation; no additional collagen or bone bioapatite samples were prepared. 81 adversely affect the results of this analysis. Bone samples for collagen extraction were

selected for each individual based on their degree of fragmentation; whenever possible,

samples were selected from bones that were already fragmented. In the case of C2-3-B and C2-3-D, no fragmented remains were available. Therefore a section approximately four centimetres long was cut using a bone saw from the distal ends of the left and right tibia, respectively, which had been broken postmortem.

Of the five mandibles available from Burial #4, only three (C2-4-A, C2-4-B and

C2-4-C) were selected for collagen analysis for comparison with the results obtained from the tooth bioapatite of these individuals. Although the recommended sample size is five grams (Knyf, personal communication, 2010), one to two gram samples of bone were obtained from the least diagnostic portion of the mandibles where they had been broken postmortem. Finally, seven right fibulae (C2-4-F1 to C2-4-F7) were identified and

sampled from Burial #4. A sample of bone four centimetres long was obtained from each fibula, preferably from sections that were broken postmortem to minimize further damage.

Although six individuals were identified in Burial #1 using the first distal phalanx of the foot, it was deemed that this bone would not have provided sufficient collagen or bioapatite for analysis. Therefore, a sample from each of the three left fifth metatarsals identified from Burial #1 was obtained. Finally, samples of bone from three right fifth metatarsals belonging to the three individuals lacking provenience were prepared for stable isotope analysis in order to compare the results with those obtained from burials of known context.

During the demineralization phase of the sample preparation, it became apparent that a number of samples would not likely yield an adequate amount of collagen for 82 analysis. As a result, a second set of samples were prepared for eight individuals (C2-3-A,

C2-3-B, C1-5-A, C2-4-F1, C2-4-F2, C2-4-F5, C2-4-F6 and C2-4-F7). These were extracted from the left tibiae of C2-3-A, C2-3-C and C1-5-A and the diaphysis of the fibulae from Burial #4 (C2-4-F1, C2-4-F2, C2-4-F5, C2-4-F6 and C2-4-F7).

Unfortunately, the bone initially sampled from C2-4-A was very poorly preserved, and a second attempt to sample collagen from this individual also resulted in problems during preparation, so neither a collagen nor a bone bioapatite sample were obtained for this individual. In total, one collagen sample each from the remaining 20 individuals was prepared for stable carbon and nitrogen isotope analysis (n = 20 collagen samples).

All bone samples were cleaned of dirt and debris (i.e., plant roots), rinsed with water, and then broken into roughly one centimetre pieces. The bone samples were cleaned ultrasonically six to eight times for 10 to 15 minutes at a time in tap water, and again three to five times for 10 to 15 minutes at a time in distilled water. The samples were then left to dry overnight. Any discolouration on the surface of the dried bones was removed using a diamond-tipped dremel (Dremel MultiPro™ Cordless, 9.6 V Model 780,

5,000-25,000 RPM). The samples were again washed ultrasonically three times for 10 to

15 minutes at a time in distilled water and left to dry for a minimum of three days in the fume hood. Once dry, the samples were weighed in order to calculate collagen yield.

The protocol used by Martin Knyf of the School of Geography and Earth Science at McMaster University, which employs a modified version of the Longin (1971) method

(Chisholm et al. 1983), was used to prepare collagen samples for stable carbon and nitrogen isotope analysis. After cleaning, two to five gram samples of bone were weighed, broken into one centimetre pieces and transferred into labelled 50 ml plastic centrifuge tubes. The tubes were then filled to the 25 ml mark with 0.25 M hydrochloric 83

acid (HCl) using a squeeze bottle. The samples were left to sit for a minimum of 20 minutes, although most were left for several hours, until few or no bubbles were released from the bone. Next, the samples were centrifuged at speed six for five minutes. After centrifuging, the HCl was decanted, fresh 0.25 M HCl was added up to the 25 ml mark and the samples were again left to sit for several hours, usually overnight. This process was repeated each day until the bone took on a soft consistency, henceforth referred to as a pseudomorph, which generally took about 4 weeks.

Once the mineral component had been removed, the pseudomorphs were centrifuged five times in distilled water to remove any residual HCl. Following this, 20ml of 0.1 M sodium hydroxide (NaOH) was added to each centrifuge tube, and the tubes were left to sit for 20 minutes in order to remove noncollagenous organic residues and humic acids. If the pseudomorphs take on a very dark (i.e., black) colour, the process is typically repeated a second time, although this step was not required in this study. The pseudomorphs were then centrifuged four times in 20 ml of distilled water at speed six for six minutes to remove the NaOH. In order to change the pH to a more acidic level in order to prevent the growth of microorganisms, 30 to 40 ml of HCl was added to each tube, the tubes were gently agitated for 2 minutes and centrifuged, and the liquid was decanted.

To remove particulate organic contaminants from the bone, the pseudomorph remaining at the bottom of each tube was transferred into a 50 ml screw cap vial, and each vial was topped up to about 40 ml with distilled water. The vials were then placed in a 1000 ml beaker partially filled with water. The beaker was placed on a hot plate set between 90°C and 100°C for six to 10 hours until the collagen was liquefied. The vials were then removed from the beakers and the contents of each were poured into their 84 corresponding cleaned centrifuge tube. The tubes were centrifuged at speed six for six minutes, to separate and remove particulate matter from the liquid. The liquid from each tube was poured into a cleaned Teflon beaker, which were placed on the hot plate until most of the liquid had evaporated. Then, the process of heating the samples in a beaker, centrifuging them, and pouring them into Teflon beakers was repeated a second time. All liquid remaining in the Teflon beakers, representing 1 to 2 ml of liquefied collagen, was transferred into pre-weighed and labelled small screw cap vials. The caps were placed loosely on the vials and they were placed in a drying oven at 37°C for seven to ten days.

Once the collagen was dried, it was weighed in order to calculate collagen yield.

5.2.2 Bone Bioapatite

In most cases, carbonate from bone bioapatite samples was extracted from the same element used for collagen extraction. No bone bioapatite was extracted from several individuals (C2-4-A, C2-4-B and C2-4-C) to minimize damage to the mandibles, and enamel bioapatite was extracted for these individuals instead. Additionally, bone from the right femur of C2-3-C was sampled for bioapatite rather than from the left tibia that was sampled for collagen extraction. It has been proposed that the isotopic composition of cortical bone, such as that recovered from long bone diaphyses, is less likely to be altered by postmortem contamination and diagenesis than that of the more porous cancellous bone (Lambert et al. 1985; Norr 1995:210). Thus samples were extracted from areas with high amounts of cortical bone whenever possible. Samples were then cleaned in the same manner as for collagen, although surfaces were thoroughly scmbbed by hand before being washed ultrasonically.

After cleaning, samples of bone were prepared using the protocol developed by 85

Martin Knyf of the School of Geography and Earth Science at McMaster University, which involves a modified version of the method outlined by Koch and colleagues

(1997). Bone samples were cmshed by hand using a mortar and pestle and approximately

100 mg of cmshed bone was weighed and placed into labelled 50 ml plastic centrifuge tubes. Following this, 4 ml of 2.5% sodium hypochlorite (NaCIO; bleach) solution was added to the tubes (approximately 0.04 ml of diluted bleach per 1 mg of sample) to remove organic material. The samples were left to react with the bleach for three days, with occasional agitation during this time.

The samples were rinsed with distilled water and centrifuged five times at speed

six for five minutes to remove the bleach. Usually at this stage 1 M acetic acid

(CH3COOH) is added to the samples; however, it has been demonstrated that the use of acetic acid to remove more soluble carbonate ions (such as calcite and absorbed carbonates) can increase crystallinity and decrease biogenic carbonate content (Garvie-

Lok et al. 2004; Koch et al. 1997; Zazzo et al. 2004). To mitigate this, after the final wash, 4 ml of 1 M acetic acid buffered with calcium acetate (Ca[C2H302J2) was added to the samples (approximately 0.04 ml of acetic acid buffer per 1 mg of sample) rather than

1 M acetic acid. The powdered bone was left to react with the buffered acetic acid for approximately 24 hours. The samples were then rinsed with distilled water and centrifuged four times at speed six for five minutes to remove the acetic acid. The samples were left in the fume hood to dry overnight, and transferred to the drying oven at

37°C for three days. Once the bone powder was dry, samples were transferred to clean two millilitre centrifuge tubes. 5.2.3 Tooth Enamel Bioapatite

Samples of bioapatite were extracted from the tooth enamel of several individuals.

Only teeth in occlusion were sampled to ensure they originated from a specific individual; thus the total number of tooth enamel bioapatite samples is six. Samples were extracted

from the left mandibular third molar of C2-3-C and C2-4-A, the right mandibular third

molar of C2-4-B and C2-4-C, the right maxillary first molar of C2-3-A, and the right

mandibular second molar from CI-5-A. These particular teeth were selected since the permanent dentition forms during childhood, and should provide a record of diet during this period.; the third molar forms roughly between the ages of eight and 15 years, and the

second molar forms between the ages of three and eight years (Moorrees et al. 1963b).

Tooth enamel bioapatite extraction was done following the protocol developed by

Martin Knyf of the School of Geography and Earth Sciences of McMaster University, based on the method outlined by Koch and colleagues (1997). Each tooth was cleaned ultrasonically twice for 10 to 15 minutes in tap water. The enamel was then wiped clean with dilute acetic acid to remove surface contaminants, and cleaned ultrasonically an

additional three times in distilled water for 10 to 15 minutes to ensure that the acid was removed. The samples were then left to dry overnight in a fume hood at room temperature.

Following cleaning, a diamond-tipped dremel (Dremel MultiPro™ Cordless, 9.6

V Model 780, 5,000-25,000 RPM) was used to clean the surface layer of enamel off the

area of the tooth to be sampled, again to eliminate surface contaminants from the sample.

Using the dremel, 20 mg samples of enamel powder were removed from each tooth and weighed into labelled 2 ml centrifuge tubes. An additional 10 to 20 mg of powder from

each sample was weighed into separate two millilitre centrifuge tubes for FTIR analysis. 87

The enamel samples for bioapatite extraction were then reacted with 0.8 ml of

2.5% bleach solution (approximately 0.04 ml of diluted bleach per mg of sample), and left to react for 24 hours with occasional agitation. After 24 hours, the samples were centrifuged at speed six for five minutes and the bleach was removed using a pipette. The samples were then washed five times in distilled water and centrifuged at speed six for five minutes each. Then, 0.8 ml of acetic acid buffered with calcium acetate was added to each sample (approximately 0.04 ml of acetic acid buffer per mg of sample) and the enamel was left to react with the acid for approximately 24 hours. After this, the samples were rinsed with distilled water and centrifuged at speed six for five minutes five times to remove the acid. The samples were then left with the lids off under the fume hood overnight, and transferred to a drying oven at 37°C to dry for approximately three days.

5.2.4 FTIR Analysis

In order to examine diagenetic alteration of carbonate in bioapatite, five bone (Cl-

5-A, C2-3-A, C2-3-C, C2-4-F4, and U-RM2) and five tooth (C1-5-A, C2-3-A, C2-4-A,

C2-4-B, and C2-4-C) samples were analyzed using FTIR. Not all samples were analyzed using FTIR, and instead a representative subsample was selected. Every attempt was made to analyze samples from each burial to test whether diagenesis was uniform across the site. Samples were prepared following the methods described by Wright and

Schwarcz (1996). Twenty milligram samples were collected from the bone cleaned for carbon isotope analysis of bioapatite prior to carbonate extraction. This clean, untreated bone was ground into a fine powder and passed through a #200 mesh sieve.

Approximately 2 mg was weighed out for each sample and ground with 200 mg of potassium bromide (KBr). The mixture was then compressed at 15,000 psi into 12 mm 88 pellets which were scanned using a Bomem Ml 00 FTIR spectrometer at Trent University to obtain absorbance spectra. The empty sample chamber was used as the background reference spectrum. After 32 scans were collected, absorbance spectra were plotted from

1800 to 400 cm"1, with a spectral resolution of 4 cm"1, and the crystallinity index was calculated for each sample (see Section 3.5.2).

5.2.5 Sample Analysis

All samples were analyzed at McMaster University under the supervision of

Martin Knyf. After preparation, 0.4 mg samples of collagen were placed along with a standard reference material into a Costech Instruments elemental combustion system

(ECS) and combusted into CO2 and N2 gases. These gases were then carried in a helium stream via a Conflo III coupling to a dual inlet continuous flow isotopic mass spectrometer (CF-IRMS). A Costech elemental analyzer (Costech ECS 4010) connected to a ThermoFinnigan DeltaPlus XP measured both the stable carbon and nitrogen isotope ratios in each collagen sample. Roughly 1.5 mg samples of bioapatite were placed into an

Isocarb with a standard where they were reacted with 100% phosphoric acid (H3PO4) at

90°C. The resulting CO gas was analyzed on a Fisson's Optima dual inlet IRMS.

5.2.6 Statistical Analysis

Many commonly utilized statistical tests include the Pearson's product-moment correlation coefficient (Pearson's r), the independent samples /-test (Student's /-test) and one-way analysis of variance (ANOVA). These approaches assume the dependent variable is measured on a continuous scale (i.e., interval or ratio data), and that the data is independent and randomly sampled. These tests also assume that the data is normally distributed and the samples are obtained from populations of equal variances (Pallant 89

2001:170-172). In order to determine whether the data from Caledonia was normally

distributed, the Shapiro-Wilk method was used (Appendix H). This method revealed that

15 13 several of the isotopic values and preservation indicators (8 N, 8 Cbi0, collagen yield, tooth enamel bioapatite CI, Early to Late Classic A Cbio-coi and Str. C-2 Tomb A Cbio-coi) were not normally distributed. Nonparametric statistical tests are more appropriate for

small samples and those that deviate from normality (Zimmerman and Zumbo 1993).

Therefore, Spearman's Rank Order Correlation (rho) was implemented to reveal whether

1 3

correlations existed between two groups (i.e. collagen yield and 8 Ccoi), the Mann-

Whitney U test was used to test for differences between two independent groups (i.e.,

13 male versus female 8 Ccoi values), and the Kruskal-Wallis test was used to test between three or more groups (i.e., 8 Ccoi values from the Late Preclassic, Early to Late Classic, and Late Classic time periods). All statistics were calculated using the statistical package

SPSS. These statistical tests are discussed in greater detail elsewhere (Madrigal 1995;

Pallant 2001:255-265; Shennan 1988; VanPool and Leonard 2011). 90

CHAPTER SIX RESULTS

6.1 Skeletal Analysis: Age, Sex, Stature and Pathology

As noted in Chapter Two, the skeletal remains from Caledonia have been previously analyzed by Dr. Herman Helmuth, and the results of his analysis can be found

elsewhere (Helmuth 1985). The first half of this chapter focuses on confirming and

expanding upon Dr. Helmuth's observations because the original detailed documentation

of this collection was not available for consultation. Any discrepancies between the original analysis (Helmuth 1985) and that presented here are attributed to the long period in which the collection was in storage, during which some accidental mixing of skeletal remains may have occurred. Here, the skeletal profiles for the individuals sampled for

stable carbon and nitrogen isotope analysis, including estimates of age and sex (Appendix

F), and pathology (Appendix E), as well as the MNI for each burial are provided.

6.1.1 Burial #1

The remains recovered from Burial #1 were the least well preserved of those examined, consisting largely of fragmented long bones as well as partial and complete bones of the hands and feet. Previous investigations of this specific burial identified seven adults and one subadult (Helmuth 1985), and, more recently, eight individuals and one subadult (Healy et al. 1998). These estimates were based primarily on the identification of the dentition, 6f which only that of the subadult was available for me to observe.

The MNI was determined based on the highest number of recurring elements. In this case, there were a total of 11 first distal phalanges of the foot, and although these elements are difficult to side, given that a single individual would have two (a left and a 91 right), a minimum of six individuals are represented. The dental remains of the subadult brought the MNI estimate to seven individuals. No pathological lesions were observed on any of the remains from this burial.

Because of the poor preservation of this burial, the age and sex of the interred individuals were difficult to assess. Despite this, the six individuals identified based on the number of first distal phalanges were probably all adult based on the size of these bones. It is also agreed that at least one individual was female because of the small size and gracility of the bones, and similarly one individual was male due to the large size of a left and right navicular bone (Helmuth 1985:418). However, because the morphological features of the skull and pelvis typically used for sex determination were not preserved, the sex of all individuals from this burial could not be confirmed. It was also not possible to age or sex the three left fifth metatarsals sampled for stable carbon and nitrogen isotope analysis in this study.

6.1.2 Burial #3

The original MNI of four individuals for this burial was confirmed. Elements were attributed to Individuals A through D based on the original investigation (Helmuth 1985).

6.1.2.1 Individual A

Individual A from Burial #3 (henceforth C2-3-A) was found to be a middle-aged probable male, at least 22 years of age at death, and likely in his mid-to-late forties (see

Appendix F). Reconstruction of the cranial vault revealed cranial modification identified as tabular erect (Imbelloni 1937), which implies that the skull was subject to fronto- occipital compression and that the pressure must have been applied to the upper squamous region of the occipital. The pressure in this form of modification involves both 92 parietals and the occipital bone (Dembo and Imbelloni 1938). This individual exhibited no dental caries or abscesses, but did have mild calculus primarily on the buccal surface of the maxillary dentition and on the lingual surface of the mandibular dentition

(Appendix E).

6.1.2.2 Individual B

Individual B (henceforth C2-3-B) was represented only by infracranial long bones. This individual was identified as an adult of unknown sex, whose age could not be accurately determined (see Appendix F). The distal epiphysis of the left humems was fused with little observable fusion line, suggesting that this individual was over the age of

10 years at the time of his/her death (Buikstra and Ubelaker 1994:43; Appendix F). This combined with the size of the bones suggests that this individual was an adult.

6.1.2.1 Individual C

A somewhat complete skeleton was identified as Individual C from Burial #3

(henceforth C2-3-C). This individual was identified as a young adult male, likely 25 years of age or younger at the time of death as the third to fifth sacral vertebrae were unfused

(see Appendix F). In addition, age estimation using the pubic symphysis provided estimates from the mid-teens to mid-twenties, confirming that C2-3-C was a young adult at the time of his death.

Reconstruction of the cranial vault revealed that this individual exhibits parallelo- fronto-occipital cranial modification (Buikstra and Ubelaker 1994:161). Although this form of cranial modification is often grouped under tabular oblique (Buikstra and

Ubelaker 1994:161; Stewart 1974), here they are considered to be two separate forms.

This is due to the fact that the parallelo-fronto-occipital modifications are formed through 93 the application of a board to the front and back of the head that are tightly bound together, while tabular oblique forms, including pseudocircular modifications, are caused by wrapping bandages around the head (Stewart 1974). These two different methods produce different appearances; the former produces bilateral parietal bulging whereas the latter produces a vertical elongation of the cranial vault. Therefore, C2-3-C exhibits parallelo- frontp-occipital cranial modification.

Mild porosity was observed on the alveolar margin of the maxilla as well as on the exterior surface of both temporal bones superior to the external auditory meatus. At this time, it is unknown what disease process may have caused these lesions. Mild porotic hyperostosis was also observed on both parietal bones along the sagittal suture as well as on the squama of the occipital bone. The lesions were healed at the time of death. In addition, an area of periosteal reaction measuring two centimetres long by half a centimetre wide was observed on the posterio-lateral surface of the distal third of the left femur. An area of periosteal reaction measuring three centimetres long by one centimetre wide was also present on the anterio-medial surface of the distal third of the left tibia.

Bilateral agenesis (hypodontia) of the mandibular lateral incisors was also observed.

Finally, two defects consistent with linear enamel hypoplasia were observed on both the left and right maxillary canines, and one defect was observed on both the left and right mandibular canines (see Appendix E).

6.1.2.4 Individual D

The final individual identified in Burial #3 (C2-3-D) was represented only by intracranial long bones (see Appendix G). The femoral head diameter indicated that this individual was a probable female (Stewart 1979:120), and epiphyseal fusion of the 94

femoral head indicates that she was over 20 years of age at the time of her death (Buikstra

and Ubelaker 1994:43; see Appendix F). No pathological lesions were observed in this

skeleton.

6.1.3 Burial #4

A minimum of seven individuals were identified in Burial #4 in this study. Only the three mandibles and seven right fibulae sampled for stable carbon and nitrogen

isotope analysis, as well as two mandibular fragments which were identified but not

discussed by the original investigator, are discussed below (Helmuth 1985)..

6.1.3.1 Individual A

Individual A from Burial #4 (henceforth, C2-4-A) was represented by the poorly preserved left half of a frontal bone as well as a similarly poorly preserved mandible. This

individual was identified as an adult of unknown age based on the emption of the third mandibular molars, and the supraorbital margin indicated that this individual was a probable male (Appendix F). Several of the teeth were broken postmortem, but two of the

crowns were recovered and refitted to the roots, which were still in occlusion. Slight

calculus was observed on the buccal surface of the right first mandibular molar, but no caries or abscesses were identified (Appendix E).

6.1.3.2 Individual B

The right half of a second mandible was identified as Individual B (henceforth,

C2-4-B). The emption of the right third mandibular molar suggests that this individual was an adult, although sex could not be determined (Appendix F). No pathology was present on the mandible or in the dentition. 95

6.1.3.3 Individual C

A second, nearly complete mandible was identified as Individual C from Burial #4

(henceforth C2-4-C). It is possible that this individual was a probable female because of her small, rounded chin. As the right third mandibular molar of this individual had empted, this individual was an adult at the time of her death (Appendix F). No skeletal pathology or dental conditions were identified.

6.1.3.4 Individuals D and E

Although the right gonial angle of a fourth individual was marked as Individual D, and the left portion of a fifth mandibular body was identified, neither were mentioned in the original report (Awe 1985; Helmuth 1985). All dentition had been lost postmortem for both mandibles, and no pathological lesions were observed. These elements were not sampled for isotopic analysis. Although these elements brought the MNI for Burial #4 to five individuals rather than the original three (see Awe 1985; Helmuth 1985), the fibulae discussed below increased this figure to seven individuals.

6.1.3.5 Fibulae

Six left and seven right fibulae were identified in Burial #4, the latter of which

(henceforth, C2-4-F1 to C2-4-F7) were sampled for stable isotope analysis. While the diaphyses of the fibulae were complete, the proximal and distal ends were broken postmortem. Neither the sex nor age estimate of the individuals they belonged to could be determined. The lengths of the diaphyses suggest that they were all adults" at the time of their death (Appendices F). After preparation of samples for isotopic analysis it was observed that the right fibula, designated C2-4-F4, exhibited woven bone formation, 96 interpreted as periostitis, on the anterior-lateral surface of nearly the entire diaphysis. No pathological lesions were observed on the remaining fibulae.

6.1.4 Burial #5

A single, relatively complete, individual (CI-5-A) was recovered from Burial #5 and was determined to be an older adult and probably female (Appendix F). Several cervical, thoracic and lumbar vertebrae, as well as the available portion of the sacrum, exhibited Schmorl's nodes and osteophytosis on their bodies. Porosity was also observed on the medial aspect of the articular surface of the left patella. The anterior surface of the midshaft of the right humems and the anterior surface of the proximal third of the right femur exhibited reactive woven bone formation. Her dentition exhibited several cases of carious lesions and antemortem tooth loss, three linear hypoplastic defects were observed on the maxillary canines, two defects were observed on the left and right lateral incisors, and one was observed on the left and right maxillary central incisors (see Appendix E).

6.1.5 Unknown Burial

A minimum of three individuals were identified from the commingled remains not associated with one of the burials. Based on the size of the bones, it is likely that all three individuals were adult, although no sex determinations could be made. No pathological lesions were identified. It is possible that these remains originated from Burial #1, but at this time, there is no way to substantiate this.

6.2 Sample Preservation

6.2.1 Preservation of Collagen

Overall, the collagen sampled from the individuals from Caledonia was well 97 preserved. The results of the methods used to assess collagen preservation are listed in

Table 6.1. Collagen yield ranged from -0.7% to 10.1% with a mean of 3.0 ± 2.4%. Of the

20 samples of collagen prepared, 17 (excluding C2-4-F1, C2-4-F5 and CI-5-1) had collagen yields above the acceptable level of 1% (van Klinken 1999). The negative value obtained for sample C2-4-F1 is likely the result of an instrumentation issue (i.e., the scale used to weigh the sample before and after treatment was not properly calibrated) or a calculation error. The atomic C/N ratios of all 20 samples ranged from 3.1 to 3.2 (mean =

3.2 ± 0.02; see Table 6.1), which fall within the accepted range of 2.9 to 3.6 (DeNiro

1985) indicating adequate sample preservation. Carbon content (wt % C) ranged from

28.0% to 40.1% with a mean of 35.5 ± 2.8%, and nitrogen content (wt % N) ranged from

Table 6.1 Indicators of Sample Preservation

Collagen Sample ID wt%C wt % N C/N Bone CI Enamel CI Yield (%) A1-1-LM1 4.5 35.5 13.1 3.2 - - A1-1-LM2 10.1 35.0 12.8 3.2 - - A1-1-LM3 2.3 37.3 13.6 3.2 - - C1-5-A 0.8 34.8 12.7 3.2 3.6 3.6 C2-3-A 3.0 28.0 10.3 3.2 3.7 3.6 C2-3-B 7.0 40.1 14.4 3.2 - - C2-3-C 2.4. 37.2 13.7 3.2 3.6 - C2-3-D 3.8 35.3 13.1 3.2 - - C2-4-A - - - - - 4.4 C2-4-B 4.5 34.4 12.5 3.2 - 3.8 C2-4-C 2.4 33.6 12.3 3.2 - 3.6 C2-4-F1 -0.7* 33.8 12.3 3.2 - - C2-4-F2 1.0 37.3 13.7 3.2 - - C2-4-F3 1.1 34.1 12.5 3.2 - - C2-4-F4 5.3 • 38.1 13.4 3.2 3.9 - C2-4-F5 '0.9 36.9 13.6 3.2 - - C2-4-F6 1.4 35.4 13.1 3.1 - - C2-4-F7 1.9 30.5 11.1 3.2 - - U-RM1 3.1 36.6 13.4 3.2 - - U-RM2 3.6 38.4 14.1 3.2 3.8 - U-RM3 2.2 38.2 14.0 3.2 - - This negative value is likely the result of an instrumentation error (i.e., the scale used to weigh the sample before and after preparation was not properly calibrated) or a calculation error. 98

10.3% to 14.4% with a mean of 13.0 ± 1.0 % (see Table 6.1). All of these values fall

within acceptable ranges (Ambrose 1990), and none were excluded from this study.

To assess whether the sample preparation techniques adequately removed

n ic

diagenetic contaminants, the correlations between 8 C or 8 N values and several preservation indicators were tested using Spearman's Rank Order Correlation Coefficient

13 15 (see Appendix H). No significant correlations were found between 8 Ccoi and 8 N (rho =

13 13 0.14, n - 20, p = 0.54), 8 Ccoi and collagen yield (rho = 0.40, n = 20, p = 0.08), 8 Ccoi

13 and wt. % C (rho = 0.39, n = 20, p = 0.09), or 8 Ccoi and C:N (rho = -0.13, n = 20, p =

0.59). Similarly, no significant correlations were found between 815N and collagen yield

(rho = 0.17, n = 20, p = 0.46), 815N and wt. % N (rho = 0.18, n = 20, p = 0.46), or 815N

and C:N ratios (rho = -0.02, n = 20, p = 0.95). The lack of statistically significant

correlations between these valuables indicates that the pretreatment procedures

adequately removed diagenic contaminants.

6.2.2 Preservation of Bioapatite

The five bone bioapatite samples analyzed using FTIR had crystallinity indices

(CI) that ranged from 3.6 to 3.9 with a mean of 3.7 ± 0.1 (see Table 6.1). These values

fall within the range of modem human bone (2.8 to 4.0; Wright and Schwarcz 1996),

indicating that the bone samples had not likely been recrystallized. The CI of the tooth

enamel samples ranged from 3.5 to 4.4 with a mean of 3.8 ± 0.3 (see Table 6.4), which are consistent with the values reported for modem tooth samples (3.8 and 4.4;

Keenleyside et al. 2011), indicating little to no diagenetic alteration.

To assess whether the pretreatment techniques successfully removed diagenetic contaminants, Spearman's Rank Order Correlation coefficient (rho) was used to 99

1 ^ determine whether a correlation existed between the 8 C values in bone and tooth and their respective crystallinity indices (see Appendix H). No significant statistical 1 -5 relationship was found between tooth enamel bioapatite 8 C values and CI (rho = 0.68, n

= 5, p = 0.20), or between bone bioapatite 8 C values and CI (rho = 0.60, n = 5, p =

0.29). Thus, the pretreatment techniques used during the extraction of bioapatite successfully removed diagenetic contaminants.

6.3 Stable Nitrogen and Carbon Isotope Data

6.3.1 Summary of Isotope Data

Tables 6.2 and 6.3 provide the details of the isotopic results. The analytical precision was 0.1 %o or better for collagen, bone bioapatite, and tooth bioapatite.

Analytical accuracy could not be calculated because the values of the lab standard were not provided. Although stable oxygen isotope data were obtained (see Appendix G), a discussion of this data was beyond the scope of this thesis. No statistically significant • correlation was found between collagen and bioapatite 8 C values (rho = -0.12, n = 18, p

= 0.98; Appendix H), which suggests but does not confirm a different source of dietary

i c IT carbon for each. No significant correlation was found between 8 N and A Cbio-coi (rho =

-0.39, n = 18, p = 0.11; Appendix H), which suggests no apparent association between protein source and level of carnivory (but see Kellner and Schoeninger 2007).

Although the collagen sample from C2-4-F4 was taken from an area of healed periostitis, according to Katzenberg and Lovell (1999:322) it is unlikely that the deposited bone is different enough to override the isotopic composition of the underlying normal bone if the lesion is small and the bone was sampled in cross section. Given that the isotopic composition of this sample is consistent with the remaining Caledonia samples

(see Table 6.2), it is likely unaffected by pathological processes.

The 813C values of the collagen samples ranged from -13.5%o to -7.4%o, with a

mean of-10.0 ± 1.9%o, indicating a strong dependence on maize-fed resources in the

11 protein component of the diet (Figure 6.1). The 8 C values of bioapatite ranged from

-8.4%o to -4.2%o (mean = -6.0 ± 1.5%o; Figure 6.2), which is consistent with substantial

consumption of maize in the whole diet.

11.0 - 10.5 - # 10.0 - • f 9.5 - • < • • £ 9.0 - • 8 5 IF - • • • • • • • • 8.0 - • 7.5 -

/.u •• i i i -14.0 -13.0 -12.0 -11.0 -10.0 -9.0 -8.0 -7.0 -6.0 13 8 Ccol(%o,VPDB)

13 15 Figure 6.1: 5 Ccoi and 8 N results for 20 sampled individuals.

• •• •• • • • • ••«••• •

-9.0 -8.0 -7.0 -6.0 -5.0 -4.0 -3.0

8»Cbio(%o,VPDB)

Figure 6.2: 5 Cbi0 results for 18 sampled individuals. 101

The 815N values ranged from 7.9%o to 10.4%o, with a mean of 8.9 ± 0.7%o (Figure

1 O

6.1), which indicates the consumption of animal protein. The A Cbio-coi values ranged

from 1.2%o to 5.9%o (mean = 4.0%o ±1.2; see Figure 6.3). Four individuals (C2-3-B, C2-3-

13 D, C2-4-F1, C2-4-F6) had A Cbi0-coi values that fell under 3%o, indicating the possibility

of marine resource consumption, three (C2-3-C, U-RM2, and U-RM3) had values within the carnivore range (3-4%o), and the remaining individuals had values that indicated

omnivorous diets. The highest spacing value (C2-4-F5, 5.9%o) was on the high end of

values reported for omnivores 1 ^

One tooth enamel sample (C2-3-A) had a 8 Cen value (-8.6%o) that was much more negative than the remaining tooth enamel samples. While the indicators suggest that

diagenesis had not occurred, this sample has been excluded from the remaining analysis because it is an outlier (see Figure 6.4). The 813C values measured in the remaining tooth

• «•• ••••» • • •

t i 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 13 A Cbi(Mol(%.,VPDB)

Figure 6.3: A Cb10-coi values for 18 sampled individuals.

o • • • * { " i i -9 -8 -7 -6 -5 -4 -3 -2 -1 0 13 8 Cen (%o, VPDB)

Figure 6.4: 8 Cen results for 6 sampled individuals (outlier sample from C2-3-A is circled). bioapatite samples ranged from -5.8%o to -2.5%o, with a mean of-4.0 ± 1.5%o (see Figure

6.4). This suggests a heavy reliance on maize in the whole diet during the period in childhood when the teeth were forming.

6.3.2 Variability by Age

The methods used to estimate the age at death of the individuals sampled are based on established standards (see Chapter 5). Due to the limited preservation of this sample, the individuals could only be assigned to the following broad age categories

(after Buikstra and Ubelaker 1994:9): young adult (20 to 35 years; n=2), middle-aged • adult (35 to 50 years; n=l), older adult (over 50 years; n=l) and adult (over 20 years; n=17). The middle-aged and older adult categories have been combined for statistical comparisons with the young adult group, and the remaining adults have been removed from the discussion. The isotopic results can be found in Tables 6.2 and 6.3, and the statistical analyses are summarized in Appendix H.

11

The collagen 8 C values for the young adult sample ranged from -10.8%o to

-8.5%o (mean = -9.6 ± 1.6%o), the values for the middle-aged adult and older adult were

-12.8%o and -11.8%o respectively, and the values for the remaining adults ranged from -

13.5%o to -7.4%o (mean = -9.7 ± 1.8%o; Figure 6.5). Although the older adults had lower

8 Ccoi values than the younger adults, this difference was not statistically significant (Z =

-0.5, p = 0.12). The young adults had bone bioapatite 813C values that ranged from -7.4%o to -5.9%o (mean = -6.7 ± l.l%o), the values for the middle- aged adult and older adult were -8.4%o and -7.7%o, respectively, and the values for the remaining adults ranged from

-8.0%o to -4.2%o (mean = -5.7 ± 1.4; Figure 6.6). Once again, there was no significant difference between younger and older adults (Z = -0.8, p = 0.12). It is also interesting that Table 6.2 Individual results of Isotopic Analysis

15 l3 Burial 8 C i 8 N 8 Cbio 8 C A r Sample ID Age Sex Time Period co en Type (%•) (%o) (%o) (%o) ^ ^bio-col A1-1-LM1 Adult - Early to Late Classic Tomb -9.1 8.2 -4.8 (%»4.4) A1-1-LM2 Adult - Early to Late Classic Tomb -8.7 8.6 -4.5 - 4.1 A1-1-LM3 Adult - Early to Late Classic Tomb -9.5 8.5 -5.3 - 4.1 C1-5-A Older Adult Probable Female Late Preclassic Simple -11.8 8.3 -7.7 -5.8 4.2 C2-3-A Middle Adult Probable Male Late Classic Cist -12.8 8.4 -8.4 -8.6* 4.4 C2-3-B Adult - Late Classic Cist -7.4 8.2 -5.0 - 2.4 C2-3-C Younger Adult Male Late Classic Cist -10.8 8.7 -7.4 -2.5 3.4 C2-3-D Younger Adult Probable Female Late Classic Cist -8.5 10.4 -5.9 - 2.6 C2-4-A Adult Probable Male Late Classic Tomb - - - -4.1 - C2-4-B Adult - Late Classic Tomb -11.7 9.7 - -5.1 - C2-4-C Adult Probable Female Late Classic Tomb -7.6 7.9 - -2.6 - C2-4-F1 Adult - Late Classic Tomb -8.9 9.2 -7.6 - 1.2 C2-4-F2 Adult - Late Classic Tomb -13.2 8.5 -8.0 - 5.2 C2-4-F3 Adult - Late Classic Tomb -11.2 8.1 -6.8 - 4.4 C2-4-F4 Adult - Late Classic Tomb -8.5 9.7 -4.2 - 4.3 C2-4-F5 Adult - Late Classic Tomb -10.3 9.0 -4.4 - 5.9 C2-4-F6 Adult - Late Classic Tomb -8.5 10.3 -6.2 - 2.3 C2-4-F7 Adult - Late Classic Tomb -13.5 8.4 -7.9 - 5.6 U-RM1 Adult - - - -10.0 9.3 -4.8 - 5.2 U-RM2 Adult - - - -8.8 9.5 -4.9 - 3.9 U-RM3 Adult - - - -8.4 8.5 -4.7 - 3.6 -10.0 8.9 -6.0 -4.0 4.0 Mean ±1.9 ±0.7 ±1.5 ±1.5 ±1.2 1;l 15 13 Note: 5 CCoi refers to collagen, 8 'C bl0 refers bone bioapatite, and 8 Cen refers to tooth enamel bioapatite Sample was removed because it was an outlier. the young adult female (C2-3-D; circled in Figure 6.6) appears to have consumed more maize than the young adult male (C2-3-C), although this difference was not tested statistically.

The young adult 815N values ranged from 8.7%o to 10.4%o (mean = 9.6 ± 1.2%o), the values for the middle and older adult were 8.4%o and 8.3%o respectively, and the values for the remaining adults ranged from 7.9%o to 10.3%o (mean = 8.9 ± 0.7%o; Figure

6.5). The older adults had slightly lower 815N values than the younger adults, but this difference was not statistically significant (Z= -0.6, p = 0.12). The bioapatite-to-collagen

11.0 -I 10.5 - * 10.0 - X X § 9.5 - < x * • Younger Adult rf 9.0 - X * •Middle-Aged Adult 8 5 sF - • X* • * x xx A Older Adult 8.0 - x XAdult 7.5 -

7.0 - 1 ! -14.0 -13.0 -12.0 -11.0 -10.0 -9.0 -8.0 -7.0 -6.0

8»Ccol(%.,VPDB)

13 15 Figure 6.5: 8 Ccoi and 8 N results by age group.

•Younger Adult • ® • •Middle-Aged Adult • A Older Adult akx X X XAdult

-9.0 -8.0 -7.0 -6.0 -5.0 -4.0 -3.0

8»Cbio(%o,VPDB)

Figure 6.6: 8 CbX0 results by age group (young adult female C2-3-D is circle). 'Table 6.3 Isotopic Summary Statistics by Sex, Age, Time Period and Burial Type U 15 U y 8 Ceol (%o) 8 N (%„) 8 8 C en (%o) A Cit)io-co l (%o) Variable Cbio (%») Mean S.D. N Mean S.D. N Mean S.D. N Mean S.D. N Mean S.D. N Sex Male -11.8 1.4 2 8.6 0.2 2 -7.9 0.7 2 -3.3 1.2 2 3.9 0.7 2 Female -9.3 2.3 3 8.9 1.3 3 -6.8 1.2 2 -4.2 2.3 2 3.4 1.1 2 Unknown -9.8 1.8 15 8.9 0.7 20 -5.7 1.4 14 -5.1 - .1 4.1 1.3 14 Age Younger Adults -9.6 1.6 2 9.6 1.2 2 -6.7 1.1 2 -2.5 1 3.0 0.6 2 Older Adults -12.3 0.7 2 8.3 0.1 2 -8.0 0.5 2 -5.8 - 1 4.3 0.2 2 Adult -9.7 1.8 16 8.9 0.7 16 -5.7 1.4 14 -3.9 1.3 3 4.0 1.3 14 Time Period Late Preclassic -11.8 - 1 8.3 - 1 -7.7 - 1 -5.8 - 1 4.2 - 1 Early to Late Classic -9.1 0.4 3 8.4 0.2 3 -4.9 0.4 3 4.2 0.1 3 Late Classic -10.2 2.1 13 9.0 0.8 13 -6.5 1.5 11 -3.6 1.3 4 3.8 1.5 11 Unknown -9.1 0.8 3 9.1 0.6 3 -4.8 0.1 3 - - - 4.2 0.8 3 Burial Type Str. A-1 Tomb -9.1 0.4 3 8.4 0.2 3 -4.9 0.4 3 0 4.2 0.1 3 Str. C-2 Tomb -10.0 2.2 9 9.0 0.8 9 -6.5 1.6 7 -3.9 1.3 3 4.1 1.7 7 Tomb Total -10.1 1.9 12 8.9 0.7 12 -6.0 1.5 10 -3.9 1.3 3 4.2 1.4 10 Cist -9.9 2.4 4 8.9 1.0 4 -6.7 1.5 4 -2.5 - 1 3.2 0.9 4 Simple -11.8 - 1 8.3 - 1 -7.7 - 1 -5.8 - 1 4.2 - 1 Unknown -9.1 0.8 3 9.1 0.6 3 -4.8 0.1 3 - - - 4.2 0.8 3

Total Mean -10.0 1.9 20 8.9 0.7 20 -6.0 1.5 18 -4.0 1.5 4 4.0 1.2 18 3 Note: N = number of individuals sampled, S.D. = standard deviation, 5' Cbio refers to bone bioapatite and 8"Cen refers to tooth enamel bioapatite.

o 106

spacing ranged from 2.6%o to 3.4%o for the young adults (mean = 3.0 ± 0.6%o), 4.4%o and

4.2%o for the middle-aged and older adult, respectively, and 1.2%o to 5.9%o (mean = 4.0 ±

1.3%o; Figure 6.7) for the remaining adults. Finally, the tooth enamel bioapatite 813C values for the young adult and older adult were -2.5%o and -5.8%o, respectively, and the values for the remaining adults ranged from -5.1%o to -2.6%o (mean = -3.9 ± 1.3%o; Figure

6.8). No statistical analyses were performed as the sample sizes were too small.

Bioapatite in tooth enamel forms during childhood and is not remodelled, whereas bone bioapatite is remodelled throughout a person's lifetime. Therefore, a comparison of the

13 13 13 8 C values in carbonate from bone (8 Ct,i0) and tooth enamel (8 Ce„) bioapatite in a

single individual can indicate if their diet changed from childhood to the last several years before death. The young adult male from the Classic period Cist burial (C2-3-C) and the older adult female from the Preclassic Simple burial (C1-5-A) were sampled for both bone and tooth enamel bioapatite. Figure 6.9 illustrates decreased maize consumption in both individuals with increased age. The decrease is much greater for C2-3-C than for

C1-5-A, although no statistical analyses were performed because of small sample sizes.

•Older Adult

•Younger Adult

AAdult •AAA AAA

I 1 1 § r 1 1 1 0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00

A"Cbfc>clU

Figure 6.7: A Cbio-coi results by age group. 107

• Younger Adult

A X X *• A Older Adult

XAdult

-7.0 -6.0 -5.0 -4.0 -3.0 -2.0 -1.0 ,3 8 Cen(%„, VPDB)

Figure 6.8: 8 Cen results by age group.

• • • C2-3-C Tooth Enamel o> • C1-5-A Tooth Enamel OC2-3-C Bone Bioapatite

s i ' DC 1 -5-A Bone Bioapatite -10 -8 -6 -4 -2 0 813C (%o, VPDB)

Figure 6.9: Difference between 8 Cbio and 8 Cen values.

Alternatively, differences in diet based on social factors (sex, social status) or time period

could be responsible for the more drastic reduction in maize consumption observed for

C2-3-C relative to C1 -5-A.

6.3.3 Variability by Sex

Three probable females (CI-5-A, C2-3-D and C2-4-C) and three probable males

(C2-3-A, C2-3-C and C2-4A) were identified in the Caledonia sample (Table 6.2).

Although it may be possible that C2-3-D was incorrectly sexed as a probable female (see

Chapter 7), here this individual is considered to be female.

1 ^

The collagen 8 C values for males ranged from -12.8%o to -10.8%o (mean = -11.8

± 1.4%o), while the female values ranged from -11.8%o to -7.6%o (mean = - 9.3 ± 2.3%o; 108

Figure 6.10). It appears that the males sampled consumed slightly less C4 protein than the

1 ^ females sampled, as indicated by their 8 Ccoi values, but this difference is not statistically significant (Z = -1.2, p = 0.25). The bone bioapatite 8 C values of males ranged from -

8.4%o to -7.4%o (mean = -7.9 ± 0.7%o) and the female values ranged from -7.7%o to -5.9%o

(mean = -6.8 ± 1.2%o; Figure 6.11). Again, it appears that the females sampled had access to a greater variety of resources based on their large standard deviations, and were relying

13 on more maize in the whole diet relative to the males sampled as indicated by their S Chi0 values , but this difference was not statistically significant (Z = -0.8, p = 0.44).

Female 815N values ranged from 7.9%o to 10.4%o (mean = 8.9 ± 1.3%o) and male values ranged from 8.4%o to 8.7%o (mean = 8.6 ± 0.2%o; Figure 6.10). Female collagen-to- bioapatite spacing values ranged from 2.6%o to 4.2%o (mean = 3.4 ±1.1 %o) and the male values ranged from 3.4%o to 4.4%o (mean = 3.9 ± 0.7%o; Figures 6.12). Neither the difference between male and female S15N values (Z = -0.6, p = 0.56), nor that of male and female A13Cbk>-coi values (Z= -0.8, p = 0.44) was statistically significant. 1 ^ Finally, the 8 C of tooth enamel bioapatite values for females ranged from -5.8%o

11.0 - 10.5 - ff 10.0 - £H 95- d 90" •Male £ 8.5 - •Female 'to A^ A Unknown 8.0 - / 7.5 -

i i S -I;5. 0 -13.0 -11.0 -9.0 -7.0 -5.0

8"Ccol(%o,VPDB)

Figure 6.10: 813Ccoi and 815N results by sex. to -2.6%o (mean = -4.2 ± 2.3%o), while the male values ranged from -4.1%o to -2.5%o

(mean = -3.3 ± 1.2%o; Figure 6.13). The difference between the male and female means

was not statistically significant (Z = -0.8, p = 0.44).

X X XMale

+ + +Female

AA A A A A A4JAA4A AUnknown

E -9.0 -8.0 -7.0 -6.0 -5.0 -4.0 -3.0 I3 8 C,si0 (%.,VPDB)

I3 Figure 6.11: 8 Cb,0 results by sex.

X X XMale

+ + +Female A A\ A AAA AAA AUnknown

I \ 5 1 1 1 ! 0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 13 A Cbi

Figure 6.12: A Cblo-coi results by sex.

XMale + A X -K +Female AUnknown 1 -7.0 -6.0 -5.0 -4.0 -3.0 -2.0 -1.0 13 8 Cenamel(%.,VPDB)

Figure 6.13: 8 Cen results by sex. 110

6.3.4 Temporal Variability

Burial #1 dated from the Early to Late Classic (Tzakol 3 to Tepeu 1) period,

Burials #3 and #4 dated to the Late Classic (Tepeu 2 and Tepeu 1 to 2, respectively),

Burial #5 dated to the Late Preclassic (Chicanel) period, and an additional three

individuals lacked both chronological and spatial provenience. The collagen 8 C value

for the Late Preclassic sample was -11.8%o, and the values for the Early to Late Classic

13 samples ranged from -9.5%o to -8.7%o (mean = -9.1 ± 0.4%o). The 8 Ccoi values for the

Late Classic period ranged from -13.5%o to -7.4%o (mean = -10.2 ± 2.1%o), and values for the unknown sample ranged from -10.0%o to -8.4%o (mean = -9.1 ± 0.8%o; Figure 6.14). It

appears that C4 protein consumption was low dor the Late Preclassic samples, was higher

in the Early to Late Classic Transition samples, and was slightly lower for the Late

Classic sample. These differences, however, are not statistically significant (x2^ 2.1, d.f.

1 ^

= 3, p = 0.56). A similar pattern is seen in the bone bioapatite 8 C values, where the value for the Late Preclassic sample was -7.7%o, and the values for the Early to Late

13 Classic ranged for -5.3%o to -4.5%o (mean = -4.9 ± 0.4%o; Figure 6.15). The S Cbio values

11.0 - 10.5 - k 10.0 - •Late Preclassic 5? 9.5- •Early to Late Classic AX ^ rf 9.0 - A Late Preclassic 85 f • A^A X Unknown 8.0 - • A • / 7.5 -

/.v { E -15.0 -13.0 -11.0 -9.0 -7.0 -5.0

8»Ccol(%o,VPDB)

13 15 Figure 6.14: 8 Ccoi and 8 N results by time period. Ill for the Late Classic samples ranged from -8.4%o to -4.2%o (mean = -6.5 ± 1.5%o), and the values for the samples of unknown provenience ranged from -4.9%o to -4.7%o (mean = -

4.8 ± 0.1%o; Figure 6.15). The difference between the 81 Cbio means by time period was not statistically significant (%2 = 5.1, d.f. = 3, p = 0.16).

The S15N value for the Late Preclassic sample was 8.3%o, the values for the Early to Late Classic period ranged from 8.2%o to 8.6%o (mean = 8.4 ± 0.4%o), the values for the

Late Classic ranged from 7.9%o to 10.4%o (mean = 9.0 ± 0.8%o) and the values for the unknown burial ranged from 8.5%o to 9.5%o (mean = 9.1 ± 0.6%o; Figure 6.14). The later samples had slightly higher 815N values than the earlier samples, but this difference is quite small (0.7%o) and statistically insignificant (x2 = 2.8, d.f. = 3, p = 0.42). The bioapatite-to-collagen spacing for the Late Preclassic sample was 4.2%o, and that for the

Early to Late Classic samples ranged from 4.1%o to 4.4%o (mean = 4.2 ± 0.1 %o; Figure

6.16). The A13Cbio-coi values during the Late Classic periods ranged from 1.2%o to 5.9%o

(mean = 3.8 ± 1.5%o), and those of the undated samples ranged from 3.6%o to 5.2%o

13 (mean 4.2 ± 0.8%o; Figure 6.16). The mean A Cbi0-coi values for the Late Preclassic and

Early to Late Classic Transition are identical, and although Late Classic values are

•Late Classic • •• •• • • • • •• + + + +Early to Late Classic A A Late Preclassic )6K XUnknown •

i -9.0 -8.0 -7.0 -6.0 -5.0 -4.0 -3.0 13 8 C,)io (%o,VPDB)

Figure 6.15: 5 Cbi0 results by time period. 112 slightly lower, the difference is not statistically significant (% = 0.02, d.f. = 3, p = 1.00).

While it appears that protein consumption was the same for all samples, extremely low collagen-to-bioapatite spacing values, which could possibly represent the consumption of marine resources, occur only in individuals living during the Late Classic period. Finally, tooth enamel stable carbon isotope ratios were only obtained for individuals from the Late

13 Preclassic and Late Classic periods (Figure 6.17). The 8 Cen value for the Late Preclassic sample was -5.8%o, and those for the Late Classic samples ranged from -5.1%o to -2.5%o

(mean = -3.6 ± 1.3; Figure 6.17); these could not be compared statistically due to small sample size.

• •Late Preclassic • Early to Late Classic • A\A • A\ • • • A Late Classic

XX X X Unknown

0.0 2.0 4.0 6.0 8.0

A»Cbi

Figure 6.16: A Cbi0-coi results by time period.

• + + -H- •Late Preclassic

i +Late Classic -8 -6 -4 -2 0

13 8 Cen(VPDB,%.)

Figure 6.17: 8 Cen results by time period. 113

6.3.5 Variability by Burial Type

As discussed in Chapter Two, three burial types are present at Caledonia - tombs, cists, and simple burials. The means values for the two tombs were not found to be

13 15 significantly different in terms of S Ccoi (Z = -0.5, p = 0.64), S N (Z = -0.6, p = 0.52),

13 13 S Cbio (Z = -0.8, p = 0.31), or A Cblo-Coi (Z = -1.0, p = 0.42; see Appendix H) values.

Thus they were combined into a single category for comparison with the other burial types (Table 6.3).

13 The S Ccoi values for the individuals interred in the tombs ranged from -13.5%o to

-7.6%o (mean = -10.1 ± 1.9%o), and the values for those interred in the cist ranged from

-12.8%o to -7.4%o (mean = -9.9 ± 2.4%o). The value for the simple burial was -11.8%o, and the values for the individuals of unknown provenience ranged from -10.0%o to -8.4%o

(mean = -9.1 ± 0.8%o; Figure 6.18). It appears that the individual in the simple burial consumed the least amount of C4 protein, followed by the tomb, cist and unknown burials, but none of these differences were statistically significant (x2 = 2.2, d.f. = 3, p =

13 0.53). The 8 Cbi'0 values for individuals interred in tombs ranged from -8.0%o to -4.2%o

11.0 -i 10.5 - I 10.0 - PS A X X4 53 9.5 - X • Simple d 9.0 - A * • Cist f 8-5 " A*a • A A A Tomb 8.0 - F X Unknown 7.5 -

1.\J 1 i t * 1 -15.0 -13.0 -11.0 -9.0 -7.0 -5.0

8"Ccol(%»,VPDB)

Figure 6.18: 8 CCoi and 8 N results by burial type. 114

(mean = -6.0 ± 1.5%o), and the values for the individuals interred in the cist burial ranged from -8.4%o to -5.0%o (mean = -6.7 ± 1.5%o; Figures 6.19). The value for the simple burial was -7.7%o, and the values for the unknown sample ranged from -4.9%o to

-4.7%o (mean = -4.8 ±0.1%o; Figure 6.19). Thus, the C4 component in the whole diet was least prevalent in the simple burial, followed by the cist, tomb and finally unknown burials, although again none of these differences were statistically significant (x2 = 1.8, d.f. = 3, p = 0.23).

The 815N values for the tombs ranged from 7.9%o to 10.3%o (mean = 8.9 ± 0.7%o), and those of the cist ranged from 8.2%o to 10.4%o with a mean of 8.9 ± 1.0%o. The value for the simple burial was 8.3%o, and those from the unknown samples ranged from 8.5%o to 9.5%o (mean = 9.1 ± 0.6%o; Figure 6.18). The nitrogen isotopic composition of individuals interred in tombs and cists is virtually identical, and no statistically significant differences were found between any of the groups (x2 = 2.2, d.f. =3, p = 0.53). The

1 ^

A Cbio-coi values for individuals buried in tombs ranged from 1.2%o to 5.9%o (mean = 4.2

± 1.4%o), and the values for those interred in cists ranged from 2.4%o to 4.4%o (mean =

13 3.2± 0.9%o). The value for the simple burial was 4.2%o, and the A Cbi0-coi for the samples

• • Simple

• • • • • Cist

AA A A A A AAA A A Tomb

X* X Unknown

-9.0 -8.0 -7.0 -6.0 -5.0 -4.0 -3.0

13 8 C bio(%o,VPDB)

Figure 6.19: 8 Cbi0 results by burial type. 115 from the unknown burial ranged from 3.6%o to 5.2%o (mean = 4.2 ± 0.8%o; Figure 6.20).

All of the groups had similar collagen-to-bioapatite spacing values, except for the cist which was slightly, but not significantly, lower than the other groups (jf = 3.8, d.f. =3, p

= 0.62).

1 ^

The 8 Cen values for the individuals buried in the tombs ranged from -5.1 %o to -

2.6%o (mean = -3.9 ± 1.3%o), the value for the cist burial was -2.5%o, the value for the individual interred in the simple burial was-5.8%o, and no tooth enamel samples were obtained from the unknown burial (see Figure 6.21). This could indicate that the individual buried in the simple burial consumed the least amount of maize in their whole diet during their childhood, followed by the individuals in the tombs and finally the

• Simple • Cist • /A AAA A Tomb XX X X Unknown

I f , , 1 ! , , 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 13 A CbioH:ol(%.,VPDB)

13 Figure 6.20: A Cb1

• Simple • A A AX XCist - ' A Tomb/Crypt

-7 -6 -5 -4 -3 -2 -1 fil3C (VPDB,%o) ° '-Tooth Enamel

Figure 6.21: 8 Cen results by burial type. 116 individual in the cist burial, but again these differences are not statistically significant (x2

= 3.2, d.f = 2, p = 0.20).

6.4 Chapter Six Summary

The age, sex, and pathology recorded for 21 individuals interred in four burials from Caledonia, Cayo District, Belize, were presented, followed by the assessment of sample preservation and isotopic results. Overall, three probable females and two probable males were identified, all sampled individuals were determined to be adults at the time of their deaths, and pathology was minimal. All samples prepared were deemed sufficiently preserved for isotopic analysis, with the exception of the tooth enamel sample from C2-3-A. Finally, the results of the stable carbon and nitrogen isotope analysis of collagen and the stable carbon isotope analysis of bone and tooth enamel bioapatite were presented by sex, age, time period, and burial type, and statistical comparisons were made. Preliminary interpretations of these results were presented, and these will be discussed in greater detail in Chapter Seven. 117

CHAPTER SEVEN

DISCUSSION AND INTERPRETATIONS

Below, the dietary interpretations gained from the results of the stable isotopic analysis are discussed. First, a general interpretation of diet at Caledonia based on the isotopic results obtained in this study is presented, followed by a discussion of dietary differences related to age, sex, social status and time period at Caledonia. The isotopic results from Caledonia are then compared with other ancient Maya sites in Belize,

Guatemala, Honduras and Mexico in order to better situate Caledonia within the Maya regional dietary context. This is followed by an examination of the relationship between pathology, health and diet at Caledonia. Finally, the chapter concludes with a review of the issues associated with isotopic analyses of diet.

7.1 Interpretation of Isotopic Analysis

7.1.1 General Diet at Caledonia, Belize

The stable carbon and nitrogen isotope values in collagen for each individual sampled from Caledonia are illustrated in Figure 7.1. As a precaution, the data in the following discussion have not been corrected to account for the fractionation between diet and tissue, unless otherwise noted, because the fractionation factors of collagen and bioapatite are not consistent between studies (Ambrose and Norr 1993; Bocherens and

Drucker 2003; Kellner and Schoeninger 2007; Tieszen and Fagre 1993; Vogel 1993).

1 ^

The calculated 8 Ccoi value for archaeological maize has been identified as -9.5%o

13 (Schwarcz et al. 1985), although in this study the compiled data provided a mean 8 Ccoi 1 % value of-8.7 ± 0.6%o (see Figure 7.1 and Appendix C). The mean uncorrected 8 Ccoi 1 ^ value for all individuals sampled in this study (-10 ± 1.9%o) approaches the 8 Ccoi values 18 -I

16 - Inshore Fish 14 -

12 - ncsnwaici risii Land Crab Offshore Fish

10 - X * X* * Reefi' S . x*x Xv* V* .« Estuarine < 8 - Terrestrial & T?:„U a* Freshwater Marine Decapods 1 ' Z 6 - 1 1 Animals to C.3 Plants Marine 4 - Invertebrates 1 2 - ^ - Maize \ Legumes Sea Grass and Algae 0 -

-2 -

_i 5 -30 -25 -20 -15 -10 -5 0

S"Ccol(%. ,VPDB)

Figure 7.1: Foodweb for Caledonia (see Section 4.3.1 for details), illustrating uncorrected human collagen values (X). 119

identified for archaeological maize, as well as the values for marine resources (means

range from -13.9 ± 3.8%o to -9.2 ± 7.0%o; see Figure 7.1 and Appendix C). These values

can be explained by the consumption of a large proportion of (a) maize; (b) maize-fed

animals; or (c) marine resources.

As discussed in Chapter 3, maize was likely the only C4 plant consistently

consumed in great enough quantity to be reflected isotopically in the bones of ancient

Maya individuals. It is most likely that the enriched 813C values of bone collagen and

bioapatite are due to the consumption of maize. The construction and maintenance, during

the Early to Terminal Classic periods, of agricultural terraces near the site to prevent soil

erosion and improve soil moisture and quality for agricultural purposes would have

allowed the Maya at Caledonia to grow a variety of crops, including maize, to support

their population (Awe 1985:32; Healy et al. 1980). Awe (1985:390) suggests that the lack

of larger, monumental buildings at the site may be attributed to the focus of labour on the

construction of these agricultural terraces. Furthermore, the mean tooth enamel bioapatite

value at Caledonia (-4.0 ± 1.5%o) is even more enriched in 13C than bone bioapatite or

collagen. Because teeth form in childhood and are not subsequently remodelled, this

suggests that childhood diets at Caledonia contained large amounts of maize. This finding

is consistent with the archaeological evidence described above.

1 "\

The 8 C values in collagen were more negative than those in bioapatite, which

could be explained by the differential fractionation that occurs between the isotopic

composition of diet and the organic and mineral portions of bone (see Section 3.4.2).

However, no statistically significant correlation was found between the two (rho = -0.12, n = 18, p > 0.05; Appendix H), which suggests, but does not confirm, a different source 1 "\ of dietary carbon for each, which is supported by the A Cbio-coi values. It is therefore 120 possible that while the carbon in bioapatite originated from dietary maize, the carbon in

collagen originated from consumed animals.

The range of 8 C values in collagen (6. l%o) was greater than the range for

bioapatite (4.2%o), which suggests greater variability in dietary protein sources,

attributable to the consumption of animal protein subsiding on a combination of C3 and

C4 plants. The mean S15N value (8.9 ± 0.7%o) indicates a substantial proportion of meat in

1 "5 the diet of the Caledonia Maya (Figure 7.1), and the mean A Cbio-coi value at Caledonia

(-4.0 ± 1.5%o) fell between the predicted values for omnivory and carnivory. The

omnivorous values are to be expected considering that the Maya are known to have

consumed a variety of plants and animals, both domestic and wild (Coe 1994; see Chapter

4), and the geographic location of Caledonia at the junction of two ecozones, as well as the nearby Macal River would have allowed the Caledonia Maya to access a variety of food resources. Three categories of animals could have contributed to the diet at

Caledonia will be discussed below, including (a) freshwater fish and mollusc; (b) terrestrial animals; and (c) marine fish and molluscs.

Freshwater fish and molluscs, including the jute sail (Pachychilus sp.) and pearly mussel (Nephronaias spp.), from the nearby Macal River could have provided a source of protein for the inhabitants of Caledonia (Awe 1985; see Healy et al. 1990; Powis 2004b).

The 8 C values of human bone collagen were more negative than those for bioapatite, which suggests the consumption of some C3 protein. This could be explained by the occasional consumption of freshwater fish, mussels, or snails. Like marine food webs, freshwater ecosystems exhibit elevated nitrogen isotope ratios because of the multiple trophic levels in their food chains (see Section 3.2.4). If the Maya at Caledonia consumed large quantities of freshwater fish, extremely enriched 815N values are expected (Wright 121

1994). Instead, the mean 815N value is closer to that expected if the Caledonia Maya consumed freshwater molluscs, which occupy a low trophic level. Species of both freshwater mussel (Nephronaias ortomanni) and snails (Pomacea and Pachyus) have been recovered from contexts at Caledonia. These species had both a dietary and ritual aspect for the ancient Maya (see Healy et al. 1990; Powis 2004b), and would have been easily collected from the nearby Macal River. The mussels were perforated, suggesting their use as ornamentation, although there is no reason to think that the flesh was not consumed prior to the shell being worked. The tips of the freshwater snail shells were nearly all broken, indicating their use as a food source (Healy et al. 1990). It is unlikely, however, that freshwater molluscs were consumed in great quantities at the site due to the fact that the collagen-to-bioapatite spacing for freshwater resources has been found to be similar to that of terrestrial herbivores (i.e., > 6.0 %o; Clementz et al. 2009). Instead, the elevated 815N mean discussed above is likely due to the consumption of terrestrial fauna at Caledonia.

The location of Caledonia at the transition between the Mountain Pine Ridge and tropical forest ecozones would have allowed the ancient Maya who lived there access to a variety of terrestrial animals, including deer, peccary, tapir, as well as a variety of reptiles, amphibians and birds, most of which would exhibit a C3 signature (see Appendix

B and Chapter Four). The only terrestrial remains recovered from Caledonia belonged to an unidentified species of deer (Awe 1985; see Section 4.2.2). Isotopic analysis of deer remains from other ancient Maya sites indicates that some animals consumed maize either through feeding/scavenging in agricultural fields or intentional feeding by the Maya

(Emery et al. 2000; van der Merwe et al. 2000; White et al. 1993:352, 2001a, 2004).

Whether intentionally or accidently, the deer consumed by the Caledonia Maya may have had access to the nearby milpas or agricultural terraces where, among other plants, maize

would have been an important crop. Although the deer bone from the site was not

analyzed isotopically, it is possible that the deer consumed by the Maya at Caledonia also

exhibited a mixed C3 and C4 S13C values due to the consumption of agricultural maize.

13 This would explain the enriched S Ccoi values of the human individuals from the site.

The 8 N, 8 Ccoi and A Cbio-coi values from the Caledonia human remains strongly

support the consumption of maize-fed terrestrial animals.

Finally, it may be possible that some of the individuals at Caledonia consumed

marine protein in large enough quantities for it to be reflected in their bone. Caledonia is

located far inland and any marine resources would have been imported to the site (Awe

1985:377). While worked and unworked marine shell from both the Caribbean and

Pacific was recovered from several Early Classic to early Late Classic contexts at

Caledonia, no marine fish remains were identified (Awe 1985; see Section 4.2.2).

Preservation is poor in the tropics, and the shelf life of marine resources would not have

been long, so it is probable that perishable marine resources were preserved using salt for

inland transport via canoe (McKillop 1995; McKillop and Sabloff 2005).

At other ancient Maya sites where an isotopic marine signature has been

identified, there is also archaeological and faunal evidence that these animals were

consumed (Powis et al. 1999; Norr 1991; White et al. 2001b; Williams et al. 2009). At

Caledonia, no marine fish remains have been recovered (Awe 1985), and while marine

shells have "been identified, they were most likely used as ornamentation or as horns, for

example, rather than as food items (Emery 2003). While marine resources were certainly traded inland to some large urban centres in limited quantities (McKillop 2010:97), the

isotopic analysis of individuals from several sites throughout the Maya region lead Gerry 123

(1997) to conclude that marine resources were unimportant foodstuffs at many inland ancient Maya sites. Finally, at Pacbitun, it is likely that marine shells were imported to the site without their animal and that marine foods were not consumed at this site (White et al. 1993). Caledonia is located near Pacbitun and the individuals from these sites share similar isotope values (see Section 7.1.6). Thus it is possible that individuals at Caledonia also imported marine shells for purposes other than food consumption, and similarly did not consume marine resources.

Four individuals (C2-3-B, C2-3-D, C2-4-F1, and C2-4-F6), had sufficiently low

8 Cbio-coi values (i.e., < 3.0%o; Ambrose and Norr 1993; Lee-Thorp et al. 1989; Norr

1995:207; White et al. 2001b) to suggest marine protein consumption. Alternatively, it may be that these low values are related to the possibility that the collagen-to-bioapatite spacing values cannot be fully explained by dietary variation (Hedges 2003; Kellner and

Schoeninger 2007). Finally, it is possible that these low A Cbio-coi values were the result of the consumption of maize-fed protein in combination with C3 plants. These four individuals have somewhat enriched 8 Ccoi values, but their 8 Cbio values fall near the average for Caledonia.

1 "\

In order to test whether the small A Cbi0-Coi values from Caledonia were related to the consumption of marine resources, the data from Caledonia were plotted according to the method outlined by both Krueger and Sullivan (1984; Figure 7.2 ), and Kellner and

Schoeninger (2007; Figure 7.3). Krueger and Sullivan (1984) provide the isotopic plots for eight theoretical diets. Figure 7.2 illustrates that while the majority of individuals from

Caledonia are located in the mixed diet, mainly maize box, one individual (C2-3-B), falls within the marine only box. 124

2 1.C3 Plants 1 3 0 2. C3 Plants + C3 Meat -1 3. C4 Plants -2 4. C4 Plants + C4 Meat -3 4 -4 5. Marine Only 5 6. Mixed, Mainly Maize 6 *(J& & - X 5 <=> "2 7. C3 Plants + Marine 8. C3 Plants + C4 Meat ^ -8 4 -9 ~-10 J-11 3 -12 <« -13 -14 -15 -16 -17 -18 -19 -20 -i 1 i 1 1 1 1 i ! 1 r T i 1 1 1 i ! ! -25 -24 -23 -22 -21 -20 -19 -18 -17 -16 -15 -14 -13 -12 -11 -10 -9 -8 -7 -6 -5 8»Ccol(%o,VPDB)

Figure 7.2: Caledonia data plotted according to Krueger and Sullivan (1984).

100% C4 Energy^ <- >M

ffl -3 O Marine o. > Protein / -5 Line

U 7b -7

100% C3 Energy -

-11 -18 -16 -14 -12 -10 8"Ccol(%.,VPDB)

Figure 7.3: Caledonia data plotted according to Kellner and Schoeninger (2007). Individuals 13 that exhibit A Cb,0-coi values below 3%o are circled. Kellner and Schoeninger (2007) argued that plotting 8 Ccoi against 8 Cbio data, and comparing these with three regression lines - one for C3, one for C4 and one for marine protein - provides a better understanding of diet. Figure 7.3 illustrates the

Caledonia data in relation to these regression lines. The majority of the individuals sampled from Caledonia fall on or near the C4 protein line, and near C4 energy, which is interpreted as indicating the consumption of maize or maize-fed protein. It is, however,

1 3 interesting that the four individuals who exhibit small A Cbio-coi values indicating marine consumption (C2-3-B, C2-3-D, C2-4-F1, and C2-4-F4), fall below the C4 protein line

(circled in Figure 7.3). Kellner and Schoeninger (2007) acknowledge that the marine protein line is problematic, and when the average values for other Maya sites are plotted, those sites that are known to have had regular access to marine resources fall in a similar area (see Figure 7.4). Therefore, it is argued here that these four individuals did consume marine resources during their life.

In summary, the Caledonia Maya relied heavily on maize as a staple in their whole diet, which was most likely supplemented with terrestrial protein and some

13 freshwater resources. The enriched 8 Ccoi values of the human remains indicate that if terrestrial fauna were being consumed, they were likely feeding on a mixed C3 and C4 plant diet. The isotopic data do not support frequent consumption of freshwater fish, although occasional consumption of freshwater mussels and snails may account for the slightly depleted carbon isotope composition of collagen relative to bone bioapatite. The

13 majority of individuals at the site exhibited A Cbi0-coi values consistent with omnivorous diets, although it is possible that four individuals may have consumed marine resources. 100% C4 Energy- _^

Legend

1. Altun Ha 2. Baking Pot Marine Protein C3 Protein Line C4 Protein 3. Barton Ramie Line ^ 4. 4< Line 5. Caledonia 6. Chau Hiix 7. Cuello 8. K'axob / -5 9. Lamanai 10. 11. Pacbitun

(O 12. San Pedro -7 13. Holmul 14. Seibal 15. 16. Copan -9 X 8 17. 18. Yaxuna 100% C3 Energy. -11 -18 -16 -14 -12 -10

8»Ccol(%.,VPDB)

13 13 Figure 7.4: Average 8 Ccoi and 8 Cbi0 values for ancient Maya sites plotted according to Kellner and Schoeninger (2007). Sites where individuals are known to have consumed fish are circled. 7.1.2 Variability by Age

Of the 21 individuals from Caledonia, only a single subadult was identified and

this individual was not sampled for stable isotope analysis. The remaining adults were

divided into four categories (see Chapter 5): younger adult (n=2), middle-aged adult

(n=l), older adult (n=T), and adult (n=17). Because of small sample sizes, the middle-

aged and older adults were combined into a single category for statistical comparison

with the two younger adults. No statistically significant differences were found between

the different adult age groups (see Appendix H), suggesting that they shared a similar

diet.

In order to investigate dietary change over the lifetime of an individual, both tooth

enamel and bone bioapatite were sampled from three individuals (C2-3-A, C2-3-C and

C1-5-A). The tooth enamel sample from C2-3-A was excluded from analysis because it

was an outlier (see Section 6.4.1). It is difficult to compare the remaining two individuals

because the sampled teeth formed at different times during childhood; enamel from the

cusp of the mandibular third molar of C2-3-C would have formed between roughly eight

and ten years of age, whereas that from the second mandibular molar of CI-5-A would have been deposited between three and a half years and six years of age (Moorrees et al.

1963b). With respect to the latter individual, it is also possible that breastfeeding may have affected the isotopic value measured in this tooth, specifically during the earlier period of formation. Although breastfeeding has been estimated to have lasted until at

least three to four years of age among the ancient Maya (White and Schwarcz 1989;

White et al. 1994, 2001b; Williams 2000; Williams et al. 2005; Wright and Schwarcz

1999), the small trophic level fractionation observed for carbon (~l%o; Schoeninger and

Moore 1992:258) will not likely affect the interpretation presented here. The tooth enamel bioapatite of C2-3-C was substantially enriched in C relative

1 "^ to that of CI-5-A, but both individuals exhibited similar 8 Ccoi values (see Figure 6.9).

The difference in enamel carbon composition between these two individuals could reflect the fact that the tooth enamel from C2-3-C reflects the diet of later childhood and early adolescence, whereas that of CI-5-A reflects the diet of early childhood. Alternatively,

C2-3-C was a male individual from a Late Classic cist burial, whereas CI-5-A was a female from a Late Preclassic simple burial. It is possible that the enriched tooth enamel bioapatite of C2-3-C reflects sex, temporal, or status differences which allowed this individual greater access to maize during childhood than CI-5-A.

In general, however, the tooth enamel bioapatite values from the Caledonia Maya were enriched in 13C relative to bone bioapatite, suggesting that children consumed more maize than adults. Metcalfe and colleagues (2009a) found a similar, albeit statistically significant, trend in tooth enamel bioapatite at Chau Hiix. Similarly, weaning onto a maize-based diet following breastfeeding was identified at Postclassic Marco Gonzalez and San Pedro in Belize (Williams et al. 2005). This inter-site comparison is based on mean isotopic values from each site, and has not been assessed by region, time period, social status or sex. Nonetheless, it tentatively appears that Maya infants were weaned onto largely maize-based diets which persisted throughout childhood. While maize remained an important dietary staple during adulthood, other plant and animal resources were also added to the diet and consumed.

7.1.3 Variability by Sex

Preservation of skeletal material in the tropics is not ideal, and many of the

skeletal traits used to determine sex were not preserved in the Caledonia collection. As a 129 result, only three probable females (CI-5-A, C2-3-D and C2-3-C) and three probable males (C2-3-A, C2-3-C and C2-4-A) were identified, and no significant sex differences in

1 3

8 C values were found. Similarly, stable nitrogen isotope analysis and collagen-to- bioapatite spacing suggest that no difference in protein consumption existed between males and females at Caledonia. The female from the Late Preclassic, however, was consuming less maize than her female counterparts during the Late Classic, and had isotopic values much closer to those of the males (see Table 6.2 in Chapter 6).

One female (C2-3-D) had higher isotopic values than nearly all other individuals

15 13 at the site. The elevated 8 N value and small A Cbi0-coi value of this individual strongly indicate marine resource consumption, which is supported by her enriched collagen

8 Ccoi value. Similarly, her 8 Cbk> (-5.9%o) was also enriched, indicating high maize consumption, although this value is not as elevated as that of other individuals at

Caledonia.

While no pathological conditions were observed in the skeleton of C2-3-D, it is possible that this individual was suffering from a disease process that was not visible on her bones. Alternatively, it is possible that elevated nitrogen isotope ratios in females instead reflect the "pregnancy effect", which involves an increase of roughly 1.0%o in

815N during pregnancy (Fuller et al. 2004). This increase is very small, however, and is likely to be obscured by dietary trends unless the woman died soon after pregnancy

(Nitsch et al. 2010). Iris thus unlikely that the elevated nitrogen isotope ratio of this individual is related to pregnancy.

Although no differences between males and females have been identified at many

Maya sites (Gerry 1993, 1997; Mansell et al. 2006; Metcalfe et al. 2009a; Tykot et al.

1996; White 1986; White and Schwarcz 1989; Wright 1997a), it appears that males consumed more maize at Pacbitun (White et al. 1993), Altun Ha (White et al. 2001b),

Copan (Reed 1994, 1999), Altar de Sacrificios, Seibal and Dos Pilas (Wright 1997a).

Furthermore, at Altun Ha (White et al. 2001b) and possibly Copan (Reed 1994) males had elevated 815N values, indicating that that they consumed greater quantities of marine resources at the former site, and maize at the latter. Isotopically, this female (C2-3-D) closely resembles males at Caledonia and at other Maya sites, suggesting that this individual may have been incorrectly sexed. Only postcranial long bones were associated with this individual and, therefore, sex was estimated using measurement of the femoral head diameter (Stewart 1979), rather than the more reliable methods involving the pelvis and cranium (see Section 5.1.2.). The femoral head diameter method was developed using a historic white population (Stewart 1979:120), which may not be applicable to a sample of prehistoric Mesoamericans. The probable female sex, however, is maintained here as both this study and the original examination by Helmuth (1985:422) are in agreement, and it is possible that this individual was simply a female who had preferential access to marine resources relative to the males interred with her (but see Section 7.1.1).

7.1.4 Temporal Variability

Since the earliest period at the site (Late Preclassic) is represented by a single individual and the majority of skeletons from the site date to the Late Classic period, it is difficult to make temporal comparisons. No significant differences were found between any isotopic measures by time period (see Appendix H). As such, it is impossible to say whether temporal changes in diet existed at Caledonia.

The Classic period "hiatus" lasted from roughly A.D. 534 to 593 and is characterized by a reduction in the dedication of stealae and monuments in the Southern 131

Maya Lowlands, particularly in the Peten region of Guatemala. Willey (1974) proposed that the hiatus was related to the decline in prosperity of elite members of society at prominent sites such as . Many archaeologists now propose that the Classic period hiatus was predominately a destruction of inscribed-monuments largely confined to Tikal

(Harrison 1999:119-124), and that warfare and conflict provide a better explanation of pauses in the production of Maya stelae rather than economics (Harrison 1999; Freidel et al. 2007:189-190). For example, at Tikal, the destruction of stelae has been linked to the defeat of this site by Caracol in A.D. 562 (A. Chase and D. Chase 1987:33, 60-61, 93).

Gaps in the chronological sequences at other sites have been found to vary in terms of

onset and duration; therefore, the implication that this "hiatus" was an event that affected all Southern Lowland sites equally is an outdated concept (see Moholy-Nagy 2003:77).

Environmental change that put pressure on the political situation has been presented as an explanation for the "hiatus" at several sites. Stalagmite evidence indicates that a drought may have occurred slightly before the Classic period "hiatus" (Webster et al. 2007), the consequences of which would have taken several decades to manifest at the cultural level. Awe (1985:393) suggests that the variety of subsistence and water resources available in the environs surrounding Caledonia, as well as the construction of agricultural terraces in the region may have allowed the site residents to flourish during this period. The Maya at Caledonia may have relied on their terraces to cultivate maize, but also other crops that supported them through a period that proved difficult for sites elsewhere in the Southern Lowlands.

In terms of ceramics, Caledonia was influenced by sites in the Peten until the end of the Tepeu 1 horizon during the Late Classic period. Internal political developments may have contributed to the "hiatus" at Tikal (Moholy-Nagy 2003) and, as a result, the 132

connections that Caledonia had with the Peten may have been strained during this period.

During the Late Classic period (Tepeu 1 to 3), the ceramics from Caledonia begin to

resemble those from sites in the Belize Valley, although Peten influence remains present

(Awe 1985:379). It is possible that the Maya at Caledonia looked to sites in Belize,

specifically Caracol, for trade opportunities and political alliances during the "hiatus",

which would explain the changes in the ceramic sequence at the site. It is possible that the

Caledonia Maya adopted a similar diet or attempted to emulate the diet at sites in Belize

with which they were allied during the'Late Classic period.

7.1.5 Variability by Socioeconomic Status

The Maya had a very complex and socially stratified society and, as in other

complex societies, food was used as a means to demarcate social boundaries (Gumerman

1997; Pohl 1994; van der Veen 2003; White 1999; Wing and Brown 1979:11-12). Those

of higher social status consumed different types or quantities of foods, although these varied from site to site. In this section, intra-site status will be discussed first, followed by

a comparison of the isotopic results from Caledonia to other ancient Maya sites.

. As discussed in Chapter Four, it is difficult to ascribe status to an ancient Maya burial based solely on the location of a burial at a site (see Becker 1992; D. Chase and A.

Chase 1992; Gillespie 2001; Pendergast 1992; White 1999). To mitigate this dilemma, researchers have utilized several mortuary variables, such as grave type, construction, burial orientation, the presence of hieroglyphic murals and burial inclusions (Gerry 1993;

Scherer et al. 2007; White et al. 1993; Wright 2006). In this study, relative status was assigned as described below.

Based on the construction type and material, the two tombs are very similar (Table 133

7.1). Both tombs were vaulted, were once plastered, and contained multiple individuals with similar burial offerings. The cist burial also contained multiple individuals, but contained fewer grave inclusions that were more ceremonial than those in the tombs

(Table 7.1). For example, the cist burial lacked the pottery, jewellery, obsidian blades and spindle whorls found in the tombs, and instead contained more ritualistic items, such as chert eccentrics and a possible ballcourt marker/altar stone were recovered (Awe 1985).

Burial #5, a simple burial, was the least elaborate at Caledonia and contained a single individual and a single ceramic body sherd (Awe 1985; Table 7.1). Not surprisingly, the grave inclusions and grave construction are linked with the type of burial and, therefore, only burial type was evaluated for determining social status.

Burial #1, the tomb in Str. A-1, most likely represents a family crypt of sequential burials (Healy et al. 1998). Similarly, it is possible that Burial #4, the C-2 tomb, represents another high ranking family at the site. Originally, Awe (1985:114-115) argued that because Burial #4 was too small to hold multiple bodies at once, two of the individuals were defleshed sacrifices (cut marks were identified on several bones from this burial although they could not be associated with an identified individual) and the third was a "noble personage". The tomb would have been too small to house the remains of the additional four individuals identified in this report as well as the original three, indicating instead that Burial #4, like Burial #1, was also a sequentially utilized family crypt. The time periods for each tomb overlap, suggesting that both tombs were active at the site during the same period. The veneration of ancestors interred within buildings at a site served ritually to name genealogical relationships and resource rights (McAnany

1995), and the re-entry of these tombs and veneration of the ancestors inside may have legitimized the power of these families at Caledonia. Table 7.1: Mortuary data for Caledonia burials*

Grave Grave Burial „ . . # Individuals Skeletal Position Grave Goods Period TyPe Construction Vaulted, >6 ceramic vessels, unworked shell, limestone 1 T t ri ' Multiple (8) Extended, east-west, head to west Tomb worked spondylus shell, jade, blocks plastered spindle whorl, mano, obsidian together 7 mini eccentrics, tapered-stemmed 2 seated, 1 extended with head to Intrusive, point, obsidian core, ballcourt 3 Late Classic Multiple (4) east, position of fourth individual Cist capstone marker/altar stone over skull of unknown extended individual >6 ceramic vessels, 1 jade bead; 3 jade/jadeite ear ornaments, 1 shell Vaulted, Extended, north-south, head to bead, 1 bone needles, 2 antler Late Classic Multiple (7) Tomb plastered floor, southwest fragments, 6 obsidian blade filled with sand fragments, 2 spindle whorls, 4 fragments of shell mosaic Late ^. . ... Flexed on left side, east-west, head Simple Simple Undecorated ceramic body sherd Preclassic ° facing west *Data from Awe (1985) 135

If social status includes a dietary component, then it is probable that the

individuals within a single tomb shared a similar diet. The isotopic measures for all three

individuals sampled from Burial #1 were very close (see Chapter 6, Table 6.1), which

suggests that diet was a factor in differentiating social status. Alternatively, the isotopic

measures in samples from Burial #4 are much more varied, but do overlap those of Burial

#1. No statistically significant differences were found between Burials 1 and 4 for any

isotopic values. These two burials were condensed into a single group (Tomb), for

comparison with the cist burial (Burial #3). The isotopic data suggests that the individuals

consumed the most maize, and were the most carnivorous at Caledonia. At other sites,

this has been associated with elevated social status (A. Chase and D. Chase 2001; A.

Chase et al. 2001; White et al. 1993, 2001b, 2006).

Burial #3 was a cist burial, most likely intrusive into the plaza floor at the base of the stairway of Str. C-2. Two of the four individuals were seated, and at least one of the

remaining individuals was extended (Awe 1985:110-111). The seated burial position is

likely an ancient Maya development and has been interpreted as a sign of importance in

Maya art (Benson 1974). Similar seated individuals have been found at

(Wright et al. 2010) and Copan (Fash et al. 2004) where the seated position was

interpreted as related to elevated social status. It is possible that the two seated individuals

interred in Burial #3 at Caledonia were elevated in social status relative to the two

extended individuals, but the identities of these individuals were not recorded.

If these individuals represent a sacrificial interment, as Awe (1985) proposed,

dietary differences between the four individuals would be expected. Similarly, if these

individuals were selected from the lower ranks of society, as was proposed for the child

sacrifices at Pacbitun (White et al. 1993), it is possible that these individuals will exhibit a different diet than those buried in the tomb. The isotopic values indicate that individuals

C2-3-A and C2-3-C consumed the least amount of maize and were the least carnivorous,

whereas C2-3-B and C2-3-D consumed substantially more maize and likely marine protein. If the consumption of maize and marine resources can be attributed to elevated

social status (Metcalfe et al. 2009a; White et al. 1993, 2001b), C2-3-B and C2-3-D were

of higher status than the other two individuals. Unfortunately, whether these individuals

were seated or extended cannot be distinguished.

The individuals in the tombs exhibited isotopic values indicating the highest level

of maize consumption, which has been associated with elevated social status at some

other ancient Maya sites (A. Chase and D. Chase 2001; A. Chase et al. 2001; White et al.

1993, 2001b, 2006). The individuals from Caledonia did not, however, exhibit the reported "palace diet" of high maize and protein consumption identified at Caracol (A.

Chase et al. 2001). Instead, it appears that the protein consumption at the site was universally high, and that it was the type of protein (C3- or C4-fed terrestrial or marine

animals) instead that varied.

Because of the small sample size, the data from Caledonia suggest but do not

confirm that individuals buried in more elaborate contexts were of higher social status than those buried in contexts that required less effort to build, such as cists or simple burials (see A. Chase et al. 2001). Alternatively, temporal and or cultural factors may

influence the relationship between diet and burial type. For example, individual CI-5-A

consumed the least amount of maize and protein in her diet relative to the other burials, but dates to the Late Preclassic. Thus, an increased reliance on maize and maize-fed

animals over time, may explain the differences between this burial and the others, rather than a difference in social status. 137

The individuals buried in the cist and tombs consumed similar types of protein,

1 3

although the smaller A Cbio-coi values of the cist individuals suggest that they were more

carnivorous, or had preferential access to marine resources relative to the individuals

interred in the tombs. It may be possible that the isotopically determined diet of the

individuals interred in Burial #3 is related to the ritualistic nature of this interment.

Finally, it should be recognized that although the burial types identified at

Caledonia have also been observed at other larger centres such as Pacbitun, where they

suggest "a substantial degree of status differentiation within the site" (White et al.

1993:350), Caledonia had been classified as a minor centre (Awe 1985). It is thus unlikely that the remains of royal members of were interred there. Rather, it

is likely that the tomb burials contained the remains of the most prominent local families at the site, whereas Burials #3 and #5 may represent dedicatory burials for the initiation of new construction phases of their associated structures. For example, the diet of the

lineage heads interred at Caledonia did not exhibit the "palace diet" identified in the royal remains from Caracol (A. Chase and D. Chase 2001; A. Chase et al. 2001) because

Caledonia was not likely a site inhabited by royal individuals. Therefore, it may be more

appropriate to combine the isotopic data into a site mean and evaluate regional

differences based on site status rather than individual status.

7.1.6 Regional Comparison of Ancient Maya Diet

Below, the mean isotopic values at Caledonia are compared to those from other

ancient Maya sites. It is important to note that this section discusses the total isotopic means for each site and, therefore, does not take into consideration variation based on

age, sex, social status, or time period. Discussions of regional dietary trends related to 138 these variables can be found in their respective sections.

13 13 The mean S Ccoi value at Caledonia (-10 ± 1.9%o) is enriched in C by nearly 2%o relative to other sites in Belize from similar time periods (Gerry 1993; Henderson 1998,

2003; Powis et al. 1999; White and Schwarcz 1989; White et al. 2001b), and sites in the

Yucatan (Mansell et al. 2006; Figure 7.5). It is probable that individuals at these sites consumed more C3-fed animals, or deriving their protein from C3 plants to a greater

11 extent than at Caledonia. The Caledonia 8 Ccoi mean is not as high as that of coastal

Belizean sites (Norr 1991; Williams 2000; Williams et al. 2009), and the enriched collagen values at these latter sites likely reflect the consumption of marine protein sources. The Caledonia mean is also less positive than inland sites in Guatemala, both highland and lowland (Gerry 1993; Whittington 2003; Wright 1994; Wright and

Schwarcz 1998), which could be explained by a higher reliance on maize or C4-derived protein resources at these sites.

13 The Caledonia 8 Ccoi mean was similar to that of Caracol (A. Chase and D. Chase

2001; A. Chase et al. 2001) and Pacbitun (White et al. 1993) in Belize, Kaminaljuyu

(Wright and Schwarcz 1998) in Guatemala, and Copan (Gerry 1993; Reed 1998) in

Honduras (see Appendix D). The similarities between Caledonia and Caracol and

Pacbitun are not surprising considering that all three sites are located in similar environments and their residents would have had access to similar animal resources. At

Kaminaljuyu and Copan it is likely that the protein found in collagen there came from maize rather than terrestrial animals (Reed 1999; Wright and Schwarcz" 1998).

13 13 The Caledonia 8 Cbi0 mean (-6.0 ± 1.5%o) is enriched in C relative to other sites in Belize including Altun Ha (White et al. 2001b), Baking Pot, Barton Ramie and Cahal

Pech (Gerry 1993), Chau Hiix (Metcalfe et al. 2009a), Cuello (Tykot et al. 1996; van der 11.0 i Site Legend 1. Altun Ha 1 X 7 Xv 10.5 - ii + 2. Baking Pot 12 • 3. Barton Ramie 4. Cahal Pech 10.0 • 5. Caledonia - 10 X 14 • 6. Caracol 23 +17 9.5 - +17 7. Chau Hiix 24+ 6X++15 +18 8. Cuello 9 X x x v- , 9. K'axob £ 9.0 - 4 2 ii'w1 10. Lamanai •^ sx l3/*j 11. Marco Gonzalez rf 8-5 - v 22 12. Moho Cay 3 X*1U Irt 13. Pacbitun 14. San Pedro «© 8.0 - 19 + • 20 — 15. 16. Altar de Sacrificios 25 A 7.5 - 17. DosPilas 27 X 18.Holmul 19. 7.0 - 26 X 20. Iximche 21. Kaminaljuyu 6.5 - 22. Piedras Negras 23. Seibal 24. Uaxactun 6.0 - i t * i t r 25. Copan -1(5. 0 -14.0 -12.0 -10.0 -8.0 -6.0 -4.0 26. Chunchucmil 8»Ccol(%o,VPDB) 27. Yaxuna

Figure 7.5: Mean 8 Ccoi and 8 N values for ancient Maya sites located in inland Belize (X), coastal Belize (•), lowland Guatemala (+), highland Guatemala (-), Honduras (A), and Mexico (x). Caledonia is circled. Merwe et al. 2000), and K'axob (Henderson 1998, 2003) (Figure 7.6). The nearby

agricultural terraces may have provided a greater and more reliable amount of accessible

maize for the residents of Caledonia in comparison to the Maya who lived at these other

sites. In contrast, the sites of San Pedro (Williams 2000; Williams et al. 2009) in Belize,

Yaxuna and Chunchucmil (Mansell et al. 2006) in Mexico, and Holmul (Gerry 1993) in

Guatemala are substantially enriched in 13C compared to Caledonia. At San Pedro this

could be the result of the consumption of a large quantity of marine resources (Williams

2000; Williams et al. 2009), although at Chunchucmil, Yaxuna and Holmul it is likely

due to reliance on maize (Gerry 1993; Mansell et al. 2006).

Not surprisingly the mean carbon isotope ratio in bone bioapatite at Caledonia is

similar to that at Pacbitun (Coyston 1995; Coyston et al. 1999), although it is also similar

to Marco Gonzalez (Williams 2000; Williams et al. 2009) in Belize, and to Seibal and

1 "\

Uaxactun (Gerry 1993) in Guatemala. The enrichment of C in bioapatite seen at Marco

Gonzalez is likely the result of marine resource consumption rather than maize or maize-

fed deer which would have to have been imported to this island site (Williams 2000;

Williams et al 2009). The enrichment seen at Seibal and Uaxactun, however, is likely

due to heavy reliance on maize (Gerry 1993).

While there are fewer studies for comparison (see Appendix D), it appears that the mean carbon isotope composition of tooth enamel bioapatite of individuals at Caledonia

(-4.0 ± 1.5%o) is higher than that at Cuello (-8.7%o; van der Merwe et al. 2000), Lamanai and Pacbitun (-5.2%o and -5.6%o, respectively; Coyston et al. 1999) in Belize, and

Chunchucmil (-6.6%o; Mansell et al. 2006) in Mexico, indicating that the children at

Caledonia consumed more maize than those who lived at these other sites. The mean 1 ^ 8 Cen value in teeth sampled from Caledonia is less positive than the mean from Iximche -3 X 18 Legend -4 + 13 1. Altun Ha 12 2. Baking Pot -5 16 3. Barton Ramie X 17 A' X 11 15 + 4. Cahal Pech 5. Caledonia • 10 6. Chau Hiix 9 X X4 14 7. Cuello 7 8. K'axob £ - i 9. Lamanai © X X 6 10. Marco Gonzalez u 3 % "8 11. Pacbitun 12. San Pedro 13. Holmul -9 H X 8 14. Seibal 15. Uaxactun X 7 -10 16. Copan 17. Chunchucmil 18. Yaxuna -11 -16 -15 -14 -13 -12 -11 -10 -9 -6 13 8 Ccol(%o,VPDB)

13 I3 Figure 7.6: Mean 8 Ccoi and S CD10 values for skeletons from ancient Maya sites located in inland Belize (X), coastal Belize (•), lowland Guatemala (+), Honduras (A), and Mexico (x). Caledonia is circled. 142

(-2.1%o; Whittington 2003) and Kaminaljuyu (-3.0%o; Wright and.Schwarcz 1998, 1999)

in the Guatemala Highlands, and Yaxuna (-2.5%o; Mansell et al. 2006) in Mexico,

indicating that the Maya who resided at these latter sites consumed even larger

proportions of maize during childhood.

The mean stable nitrogen isotope value at Caledonia (8.9 ± 0.7%o) was similar to

that found at other sites including Baking Pot (Gerry 1993), Barton Ramie (Gerry 1993),

Cahal Pech (Gerry 1993; Powis et al. 1999), Cuello (Tykot et al. 1996; van der Merwe et

al. 2000), and Pacbitun (White et al. 1993) in Belize, and Holmul (Gerry 1993), Piedras

Negras (Scherer et al. 2007) and Kaminaljuyu (Wright and Schwarcz 1998) in

Guatemala. Excluding Kaminaljuyu, all of these sites are either located on the banks, or

within several kilometres, of rivers. As at Caledonia, it is possible that these nearby water

sources provided the ancient Maya at these sites access to freshwater molluscs and/or

game animals such as deer. It is also possible that because of the small size of Caledonia,

overexploitation of terrestrial animal resources was minimal, and thus these species were

available in greater quantities for consumption than at other sites such as Copan (Gerry

1993; see below).

The mean stable nitrogen isotope ratios from individuals interred at Seibal (Gerry

1993; Wright 1994, 1997a, 2006) in Guatemala, and Altun Ha (White et al. 2001b),

K'axob (Henderson 1998, 2003), Lamanai (White and Schwarcz 1989), Moho Cay (Non-

1991), Marco Gonzalez, and San Pedro (Williams 2000; Williams et al. 2009) in Belize

are higher than the mean value from Caledonia (Figure 7.3). The greater consumption of

marine resources at the Belizean sites, and the increased consumption of terrestrial

protein from higher trophic levels (i.e., omnivorous dogs or carnivorous jaguars) at Seibal

would account for the elevated S15N values at these sites. In contrast, the mean 815N value 143

at Caledonia is higher than that at Altar de Sacrificios, Itzan (Wright 1994, 1997a, 2006), and Iximche (Whittington 2003) in Guatemala, Copan (Gerry 1993; Reed 1998) in

Honduras, and Yaxuna and Chunchucmil (Mansell et al. 2006) in Mexico. The most probable explanation for this difference is the consumption of freshwater resources at

Caledonia. Alternatively, low 8 5N values were attributed to the consumption of nitrogen- fixing legumes at Copan (Reed 1998), and the low nitrogen isotope ratio at the other sites may be related to this as well.

13 13 The values illustrated in Figure 7.6 represent the mean 8 Q,io and S Ccoi values for individuals from all time periods, and of all ages, sexes and social statuses at each site,

some of which were calculated by the present author (see Appendix D). As expected, individuals interred at inland Belize sites have lower 8 Cbio and 8 Ccoi values and tend to cluster with one another, while those from sites in Guatemala (excluding Seibal; Gerry

13 13 1993) and Copan in Honduras, demonstrate slightly higher S Cbi0 and 8 Ccoi values

(Figure 7.4). The individuals sampled from Chunchucmil and Yaxuna in Mexico have the

13 13 most enriched 8 Cbi0 values and the most depleted 8 Ccoi values, indicating low meat consumption (Mansell et al. 2006), whereas those from Marco Gonzalez and San Pedro in coastal Belize demonstrate the opposite, indicating the consumption of marine resources

(Williams 2000; Williams et al. 2009). The values from Caledonia and nearby Pacbitun cluster with sites from Guatemala rather than with other sites in Belize (Figure 7.6).

Caledonia, for example, is quite close to the A Cbio-coi value from Seibal, which likely reflects the fact that both sites are located in a riverine setting and would have had access to similar animal resources.

Several isotopic studies of ancient Maya skeletal remains have published the

13 A Cbio-coi values (Gerry 1993; Mansell et al. 2006; Metcalfe et al. 2009a; White et al. 2001b; Williams 2000; Williams et al. 2009), but many have not. In cases where the

means of both the collagen and bioapatite stable carbon isotope ratios from a site were

published (Coyston et al. 1999; Henderson 1998, 2003; Tykot et al. 1996; van der Merwe

13 et al. 2000), the A Cbi0-coi was calculated by the present author (see Appendix D).

I3 The mean A Cbi0-coi value of individuals interred at Caledonia (4.0 ± 1.2%o) is

similar to that from Altun Ha (White et al. 2001b), Baking Pot, Barton Ramie, and Cahal

Pech (Gerry 1993) in Belize. These sites are all located along the Macal and Belize

Rivers, and access to freshwater resources and possibly to terrestrial animals that use the

rivers as a water source, may explain their similar collagen-to-bioapatite values. The

mean collagen-to-bioapatite spacing at Caledonia is also similar to the mean at Pacbitun

(Coyston et al. 1999; White et al. 1993), which is not surprising considering that the

residents of these sites had access to and likely consumed similar food resources. Finally,

the Caledonia mean is also similar to that from Seibal and Uaxactun (Gerry 1993) in

Guatemala and to Yaxuna (Mansell et al. 2006) in Mexico. It is possible that terrestrial

meat consumption at these sites was similar in quantity and quality to that at Caledonia.

1 ^

The spacing between bone bioapatite and collagen 8 C values reported at Marco

Gonzalez and San Pedro in coastal Belize during the Postclassic and Historic periods

(Williams 2000; Williams et al. 2009), as well as Chau Hiix during the Early Classic,

Postclassic and Historic periods (Metcalfe et al. 2009a), was much lower than the mean

spacing at Caledonia during the Late Preclassic and Classic periods (see Appendix D),

indicating that the individuals interred at these sites consumed substantially more marine

resources than those at Caledonia because they were located closer to the coast. The mean value at Cuello during the Preclassic (Tykot et al. 1996; van der Merwe et al. 2000; see

Appendix D) was also lower than that at Caledonia, although Cuello is located much 145 closer to the Caribbean coast and the Maya there would likely have had access to marine protein more frequently or in larger quantities than at Caledonia. In contrast, the

Caledonia mean was lower than that at Holmul (Gerry 1993) in Guatemala, Copan (Gerry

1993) in Honduras, and K'axob (Henderson 1998, 2003) and Lamanai (Coyston et al.

1999; White and Schwarcz 1989) in Belize. The consumption of freshwater resources at

Lamanai may explain this elevated spacing, whereas at Copan (Reed 1994) and possibly at Holmul it is probable that meat consumption was quite low. To summarize, while the collagen-to-bioapatite spacing values of four individuals at Caledonia suggest they may have consumed marine resources frequently enough and in large enough quantities for the signature to appear in their skeletons, the majority of individuals at the site did not consume marine resources to any significant degree.

In general, the isotopic values from individuals interred at Caledonia are somewhat different from those from other sites in Belize, excluding nearby Caracol and

Pacbitun. Instead, the mean isotope values are much closer to those found at several sites in the Pasion Valley region of Guatemala. The archaeological evidence suggests that the residents of Caledonia were influenced by those in the Peten of Guatemala until the end of the Tepeu 1 (early Late Classic) horizon. Following this, the Peten influence remains visible, although influences from the Belize Valley are increasingly evident than in earlier periods (Awe 1985:379). By the end of the Tepeu 3 (late Late Classic) horizon, architectural and ceramic evidence suggests that Caledonia was trading with other Peten sites, perhaps through individuals who lived at Caracol (Awe 1985:398-399).

Caledonia has been described as a minor centre, which indicates that politically

"the site came under the influence, if not actual control, of a larger primary centre" (Awe

1985:387). The nearest primary centre is Caracol, and it is likely that Caledonia was 146 subservient to this site (Awe 1985:388). For example, it could be that the Caledonia Maya were collecting granite, pine and mussel shells for export, and individuals at Caracol controlled these items and acted as liaisons for these transactions, although it is uncertain whether the relationship between Caracol and Caledonia was economic, political, or both.

A similar relationship was observed between the minor centre Chau Hiix and the primary centres of Altun Ha during the Late and Terminal Classic, and Lamanai during the

Postclassic periods (Metcalfe et al. 2009a), and all three sites shared similar diets.

Perhaps the Maya at Caledonia were emulating not only the material culture of their

Caracol and Peten allies/sovereigns, but also their diet.

7.2 Health, Pathology and Isotope Analysis

Many of the pathological conditions present in ancient Maya skeletons have been related to maize-based diets because of the nutritional deficiencies of maize itself. Maize is deficient in the essential amino acids lysine and tryptophan, and in niacin, a member of the vitamin B complex. Maize is also low in iron and contains phytates that can inhibit the absorption of this element (Huss-Ashmore et al. 1982:417). Alkali cooking, however, enhances the balance of essential amino acids and frees otherwise nearly unavailable niacin. Therefore, without the use of specific preparation techniques, the nutritional value of maize as a dietary source "is at best marginal, and any human population that attempted to depend on it as a major staple would suffer some degree of malnutrition"

(Katz et al. 1974:773). Ethnographic accounts of the Lacandon Maya of eastern-Chiapas,

Mexico, explain that after soaking maize kernels in water containing burned limestone or ground and slaked snail shells, it was ground with a metate and mano (Nations 1979). It is expected that the ancient Maya also added lime to their maize before consumption (Katz 147

etal. 1974; White and Schwarcz 1989:466).

Among the historic Maya, maize was an important staple and was consumed at

every meal in one form or another (Tozzer 1941). The importance of maize in earlier

periods has also been substantiated by palaeobotanical, archaeological and isotopic

studies (see Section 4.1.1). Although it is now recognized that the ancient Maya utilized a

variety of floral and faunal resources, in addition to maize, a maize-based diet remains the

primary suspect for iron deficiency anemia as evidenced by the presence of porotic

hyperostosis in Maya skeletal remains (D. Chase 1997; Scherer et al. 2007; Whittington

2003; Whittington and Reed 1997; Wright and Chew 1998; White et al. 2006a; but see

Williams et al. 2009; Wright and White 1996; but see Walker et al. 2009). Furthermore,

the high carbohydrate content of a maize-based diet has been linked to a high prevalence

of dental disease (Evans 1973; Huss-Ashmore et al. 1982; Larsen 1987; Ortner 2003;

Powell 1985; Tayles et al. 2000; Whittington 1999). Below, both of these conditions

within the Caledonia sample are compared to the isotopic results, and the consequences of

a maize-based diet are critically evaluated.

7.2.1 Porotic Hyperostosis

As discussed briefly in Section 4.1.3, porotic hyperostosis is a condition

characterized by the expansion of the diploe and the development of porous lesions on the

external cranial vault that can result from a variety of metabolic disorders, including iron

deficiency anemia. Although originally thought to result from iron-deficient diets, such as those based on maize (D. Chase 1997; Saul 1977; Scherer etal. 2007; Whittington 2003;

Whittington and Reed 1997; Wright and Chew 1998; White et al. 2006a; but see

Williams et al. 2009; Wright and White 1996), porotic hyperostosis has also been linked 148 to chronic parasitic infection (Holland and O'Brien 1997) and megaloblastic anemia

(Walkeretal. 2009).

Some researchers argue that gastrointestinal parasites, such as whip worm

(Trichuris) and hook worms (Necator and Anaylostoma), cause increased intestinal bleeding, and other parasites such as giardia (Giardia lamblia) and roundworm (Ascaris lumbricoides) inhibit iron absorption, both of which would lead to iron deficiency anemia

(Wright and Chew 1998). Others suggest that several species of gastrointestinal parasites can deplete the body of vitamin Bn (Walker et al. 2009:114-115). In individuals who suffer from malabsorption and chronic diarrhea, low B12 intake can lead to severe megaloblastic anemia (Walker et al. 2009:115) rather than iron deficiency anemia.

A regional comparison of porotic hyperostosis frequencies throughout the Maya region has been summarized elsewhere (Scherer et al. 2007:93-95; Wright and Chew

1998; Wright and White 1996:157-162), and indicates that this condition was more common in sites in the Peten during the Classic period than in Belize or the Yucatan during the Postclassic. Frequencies of porotic hyperostosis among the Maya are relatively high, but do vary spatially and temporally (see Wright and Chew 1998). For example, 60-

90% of adult crania from Classic period inland sites exhibit healed porotic hyperostosis

(Wright 1997b; Wright and Chew 1998:927). The lower frequency at Preclassic Cuello

(12.5% for subadults and 3.6% for adults; Saul and Saul 1991) has been attributed to less maize consumption as indicated by isotopic analysis (Wright and White 1996), while the increase in frequency of porotic hyperostosis at Lamanai from the Postclassic (1% of juveniles and 4% of adults) to Historic periods (7% of juveniles and 13% of adults) has been attributed to increased maize consumption and the introduction of Spanish infectious disease (White et al. 1994). The consumption of large quantities of maize, however, is not always associated with high frequencies of porotic hyperostosis in Maya skeletal

assemblages. For example, the low prevalence of this condition at Preclassic Iximche is

associated with high levels of maize consumption indicated by isotopic analysis

(Whittington 2003). Similarly, at Caracol, where a diet rich in maize was identified

isotopically (A. Chase and D. Chase 2001; A. Chase et al. 2001), very few individuals

(n=8) exhibited porotic hyperostosis (D. Chase 1994, 1997). Finally, parasitic infection resulting from the consumption of marine fish was proposed as a factor that may have

contributed to the expression of porotic hyperostosis at Marco Gonzalez and San Pedro, although the frequency of this condition was low (12% for both sites combined; White et al. 2006a).

Of the 21 individuals identified in the Caledonia sample, three had crania that were complete enough to assess the presence of porotic hyperostosis. None of the individuals exhibited cribra orbitalia, and two individuals (C2-3-A and C2-3-C) exhibited healed porotic hyperostosis (see Sections 6.1.2.1 and 6.1.2.3), indicating that both individuals were afflicted with the condition during childhood and that it had subsequently healed. As all three individuals were adults, the isotopic values measured in their bones reflect the isotopic composition of their diet during adulthood, not their

1 "\ childhood. Thus, the 8 Cen values are most useful here because they preserve the dietary signature during the period in childhood when the porotic lesions were active. The

13 individual C2-3-C who exhibited porotic hyperostosis had a less negative 8 Cen value than individual CI-5-A, indicating that he consumed substantially more maize than Cl-5-

A during childhood (Table 7.2). It is therefore possible that the high proportion of maize in the diet of C2-3-C during childhood inhibited iron uptake in the body, which led to anemia and eventually porotic hyperostosis. Table 7.2: Data for individuals with and without porotic hyperostosis

13 15 13 13 13 TnHiviHll . Time Burial Porotic S Ccol 5 N 8 Cbio A Cbi„col 8 Ceil Period 8 Type Hyperostosis (%o) (%o) (%o) (%o) (%o) - . . Late Middle r P. Male Cist Yes -12.8 8.4 -8.4 4.4 - Classic Adult

Late C2-3-C Young P. Male Cist • Yes -10.8 8.7 -7.4 3.4 -2.5 Classic Adult T ate C1-5-A _ , . Older Adult P. FemalFemalee Simple No -11.8 8.3 -7.7 4.2 -5.8 Preclassic r 151

While parasitic infection may also have been a contributing factor to the

development of porotic hyperostosis in the crania from Caledonia, this is unlikely. First, the risk of contracting a parasite is related to the increased waste and decline in sanitation

associated with a large population, which Caledonia did not have. Second, the contamination of water with parasites was not likely an issue considering that the likely

source of water was the relatively fast-flowing and perennial Macal River, which would have reduced the chance of parasitic infection (see Walker et al. 2009). Finally, it is unlikely that parasitic infection from the consumption of marine fish contributed to the development of porotic hyperostosis at this site because the isotopic values of the three individuals who were represented by relatively complete crania do not indicate that they consumed marine resources (see Table 7.2).

Alternatively, the porotic lesions observed on the crania of two individuals at

Caledonia may be the result of megaloblastic anemia, which is characterized by abnormally large red blood cells and is usually caused by a combination of folic acid and vitamin Bn deficiency (Herbert and Zalusky 1962). This type of anemia cannot be detected isotopically and, therefore, cannot be positively identified in the Caledonia skeletons.

In summary, it is possible that high levels of maize consumption were responsible for the expression of porotic hyperostosis. However, the possibility that either parasitic infection or megaloblastic anemia contributed to the manifestation of porotic hyperostosis in the skeletons from Caledonia cannot be eliminated. " •

7.2.2 Dental Pathology

In areas of the world where bone does not preserve well, such as the tropics of 152

Central America, teeth are often the only human tissue that survives. As such, this tissue

has been analyzed in many studies of ancient Maya remains (D. Chase 1994; Evans 1973;

Saul 1977; Magennis 1999; White 1988; White and Schwarcz 1989; Whittington 1999).

Below, several conditions of the dentition identified at Caledonia are discussed in terms

of the isotopic data. The dentition from Burial #1 was not available for observation, but

all other available dentition from Caledonia was examined. The discussion below focuses

on the dentition from individuals sampled for isotopic analysis, unless otherwise noted.

7.2.2.1 Linear Enamel Hypoplasia

As explained in Section 4.1.3, linear enamel hypoplasia is characterized by pits or

grooves in the enamel caused by a period of stress related to a variety of factors including

disease and nutrition during the time the tooth was forming in childhood. As with porotic hyperostosis, the expression of linear enamel hypoplasia in ancient Maya skeletons is

relatively common but the prevalence varies somewhat by site. For example, the prevalence of linear enamel hypoplasia on the mandibular canine of individuals interred

at several sites in the Pasion Valley was 59% (Wright 1997b), whereas at Caracol, the prevalence was only 16% of the total burial sample (D. Chase 1994, 1997). Overall, it

appears that the frequency of linear enamel hypoplasia in the dentition from Caledonia

(see Appendix E) is lower than that at other ancient Maya sites, but similar to that from nearby Caracol.

At Caledonia, two teeth from the loose dentition from Burial #3 exhibited linear enamel hypoplasia, but they could not be associated with one of the identified individuals.

Loose dentition from Burial #4 was examined and no enamel defects were observed. The teeth that typically exhibit linear enamel hypoplasia (i.e., incisors and canines) were not 153 available for the three mandibles from Burial #4, and the molars that were present did not exhibit this condition. Of the remaining three individuals, CI-5-A and C2-3-C exhibited linear enamel hypoplasia on more than one tooth, whereas C2-3-A exhibited none

(Appendix E). The childhood diet of these individuals was assessed through the analysis of stable carbon isotopes in their tooth enamel. The portion of the mandibular second molar sampled for CI-5-A would have formed between three and a half years and six years of age, and would contain carbon from the latter half of the period when the enamel defects would have formed in the incisors and canines. Unfortunately, the portion of the mandibular third molar sampled for C2-3-C would have begun to form later in childhood, between the ages of seven and a half and 13.5 years, and would not contain carbon from the period in life when enamel hypoplastic defects were formed. For comparative purposes, it is assumed here that children maintain a constant diet until adulthood, as implied above (Section 7.1.2). The isotopic analysis of the dental enamel revealed that

CI-5-A was consuming less maize during childhood but exhibited more numerous enamel defects on more teeth than C2-3-C. Although tentative, this suggests that a maize- based childhood diet did not directly result in the expression of linear enamel defects in the individuals from Caledonia, and other factors were involved.

7.2.2.2 Dental Calculus

Dental calculus is characterized by calcified plaque on the enamel surface of the dentition (see Chapter 4). A diet high in protein can result in a significant degree of calculus deposition on the surfaces of teeth (HiUson 1979). Healy and colleagues (1998) found heavy dental calculus build up on teeth, especially the molars, from Burial #1 at

Caledonia. Similar deposits of calculus were observed on the dentition from individuals If IT

interred in Burials #3 and #5 (Appendix E). The average 8 N and A Cbio-coi values from

the individuals sampled from Caledonia support a diet that includes regular protein

consumption.

The deposition of calculus on the dentition from Caledonia is consistent with

studies of dental material from other Maya sites (Evans 1973; Saul 1972). It is interesting,

however, that only 2.1% (n = 7) of individuals at the nearby site of Caracol exhibited

moderate to severe calculus, where 815N values suggest that individuals consumed more

protein (D. Chase 1997). This discrepancy may be explained by the different methods

employed to describe and quantify calculus in these different studies.

7.2.2.3 Dental Caries

As discussed in Section 4.1.3, dental caries has been positively correlated with

dietary carbohydrates, although the amount of carbohydrates is not as important as the

consistency and frequency of their consumption (Bibby 1961; Powell 1985). Scherer and

colleagues (2007:95-96) found that when compared regionally, inland sites such as Copan

and those in the Peten exhibit higher frequencies of carious lesions per teeth than sites in

Belize or the Yucatan, which they attributed to a greater reliance on maize at the former

sites. In contrast, Seidemann and McKillop (2007) found that although the frequency of

caries varied temporally and spatially among the ancient Maya, these patterns are site-

specific, rather than reliant on region (see Table 7.3). The discrepancy between these two

studies could be related to the various methods used by different researchers; some researchers provide the percentage of affected individuals (e.g. D. Chase 1994, 1997;

Scherer et al. 2007), while others provide the frequency for all the observed teeth (e.g.,

Cucina and Tiesler 2003; Cucina et al. 2003), which complicates comparisons of results from different studies. In addition, the manner in which carious lesions were scored can also influence frequency results, and therefore the comparability of different studies.

Regardless of how caries were classified or quantified, all studies propose that the increased prevalence of caries is related to elevated consumption of maize (Scherer 2007;

Seidemann and McKillop 2007).

The frequency of caries at Caledonia (Table 7.3) was calculated using teeth that exhibited at least one carious lesion from all available teeth, regardless of whether they were in occlusion, as did all other studies listed in Table 7.3. Only one individual from

Caledonia (CI-5-A) exhibited dental caries, and it is interesting that that individual also likely consumed the least amount of maize, as indicated by her 813C values (see Chapter

6). Dental caries is an age-dependent condition and the presence of this condition increases with age (Carlos and Gittelsohn 1965). Similarly, women are at a higher risk for

Table 7.3: Caries Frequencies by Tooth Count from Selected Ancient Maya Sites

Site Location Time Period Frequency (%) Source Caledonia Inland Belize Late Preclassic 13.3 This Study Late Classic 12.7 Whittington Copan Inland Honduras Classic 17.9 1999 Cuello Inland Belize Preclassic 36.9 Saul and Saul 1997 Highland Whittington Iximche Late Postclassic 8.5 Guatemala 2003 Protoclassic 14.3 Early Classic Kichpanha Inland Belize 11.1 Magennis 1999 Late/Terminal 28.5 Classic Late Preclassic 20.0 White 1986; Early Classic 24.0 Lamanai Inland Belize White and Late Classic 18.0 Schwarcz 1989 Terminal Classic 1.8 Inland Belize Late Classic 40.0 Saul 1975 Wild Cane Costal Belize Postclassic 36.2 Seidemann and Cay McKillop 2007 Note: Frequency is calculated by tooth count for all studies. dental canes due to a number of factors, including sex hormones, charactenstics of the

saliva and factors during pregnancy (see Lukacs 2008), and caries frequencies appear to

be higher among ancient Maya females (Cucina and Tiesler 2003; Cucina et al. 2003). At

Caledonia, CI-5-A was a female over the age of 40 years at the time of her death, which

may better explain her oral health than a maize-dependent diet. It may be that the low

frequency of caries at this site is due to the age and sex distribution of the sample (i.e.,

many of the individuals were younger adult males; see Chapter 6) rather than diet.

7.2.2.4 Dental Wear

Nearly all of the dentition in the Caledonia sample exhibited at least minimal

dental wear (see Appendix E). Dental wear is an age-related condition, and the degree

increases with increasing age. This condition can also remove the pits and fissures of

molar cusps where carious lesions can form, resulting in lower rates of caries. Among the

ancient Maya dental wear can also be attributed to the processing of plant foods with

manos and metates made of stone (see Chapter 4). It is possible that Caledonia partially

controlled the trade of granite for the production of manos and metates, and several of

these tools have been recovered from different contexts at the site (Awe 1985). Although the presence of dental wear in the dentition from Caledonia cannot tell us what types of

foods were consumed, it does suggest the use of manos and metates to process plant

foods. Recent research has revealed that manos and metates were used to process a variety of plant species (Lentz et al. 1996), and the isotopic data from Caledonia has

determined that while maize was a dietary staple, other plant foods were also important in the diet of the Caledonia Maya. Furthermore, the moderate degree of dental wear in the

dentition from Caledonia may explain the low rate of caries observed in this sample compared to caries rates in dental samples from other ancient Maya sites.

7.2.3 Pathological Summary

In ancient Maya studies there has been a tendency to associate a maize-based diet with poor health. Although isotopically maize was the primary staple in their diet, a variety of food resources were available to and consumed by the Caledonia Maya.

Comparisons of the presence of pathological conditions and isotopic results of the sampled individuals are inconclusive; while the presence of porotic hyperostosis and dental calculus in the skeletons from Caledonia may be related to substantial maize consumption, other conditions such as dental caries, linear enamel hypoplasia and dental wear cannot be directly related to a maize-based diet. However, the sample size is quite small and these conclusions must therefore be considered tentative.

7. 3 Methodological and Theoretical Concerns

The results discussed above are interesting, but there are several concerns that must be addressed in order to substantiate these findings. The effects of sample size, the representativeness of the sample, and the comparability of results between studies will be discussed below.

7.3.1 Sample Size

While the preservation of human skeletal and dental remains at Caledonia was very good relative to other ancient Maya skeletal collections (see below), the sample size was small. Skeletal collections from ancient Maya sites tend to be small due to issues of preservation, burial practices and excavation strategies (see following sections). Table 7.4 illustrates the combined number of individuals sampled for different isotopic analyses 15 13 Site 6 C i 8 N CMO A r Source co 8 ** '-'bio-col 8 Cen Belize Altun Ha 44 44 47 43 - White etal. (2001b) Baking Pot 9 9 4 4 - Gerry (1993) Barton Ramie 7 7 6 6 - Gerry (1993) Cahal Pech 11 11 - - - Powis etal. (1999) Caledonia 20 20 18 18 5 This study Caracol 85 85 - - - Chase and Chase (2001), Chase et al. (2001) Chau Hiix 28 28 28 28 80 Metcalfe et al. (2009a) Cuello 28 23 16 - 33 Tykot et al. (1996), van der Merwe et al. (2000) K'axob 10 11 23 - 1 Henderson (1998, 2003) Lamanai 51 49 42 - - Coyston et al. (1999), White and Schwarcz (1989) Marco Gonzalez 37 37 37 37 - Williams et al. (2009) Moho Cay 8 8 - - - Norr (1991) Pacbitun 20 20 18 - 3 Coyston et al. (1999), White et al. (1993) San Pedro 29 29 29 29 - Williams et al. (2009) Guatemala Aguateca 8 7 - - - Wright (1994) Altar de Sacrificios 54 53 - - - Wright (1994) Dos Pilas 19 19 - -• - Wright (1994) Holmul 14 15 22 13 - Gerry (1993) Itzan 5 5 - - - Wright (1994) Iximche 13 13 - 43 - Whittington (2003) Kaminaljuyu 24 15 - - 96 Wright and Schwarcz (1998) Piedras Negras 45 45 - - - Scherer et al. (2007) Seibal 47 60 27 27 - Gerry (1993), Wright (1994) Uaxactun 6 5 2 1 - Gerry (1993) Honduras Copan 46 46 36 32 - Gerry (1993), Reed (1998) Mexico Chunchucmil 3 3 5 3 3 Mansell et al. (2006) Yaxuna 3 3 19 3 14 • Mansell et al. (2006) 159

from ancient Maya sites. Not surprisingly, at major sites such as Altar de Sacrificios and

Caracol, where archaeological investigations have been ongoing for decades, the sample

sizes are quite large. In contrast, smaller sites that are often excavated for a single season, as was the case at Caledonia, yield small skeletal samples. Sample sizes are further reduced if samples are divided into subgroups based on age, sex, social status, and/or time period.

These small sample sizes are problematic for two reasons. First, it is difficult to make generalizations about the population at a site based on only a few individuals who lived there. Second, small sample sizes can obscure or exaggerate statistical trends in isotopic data. Although no precise sample size has been recommended for comparisons of isotopic data, in order for a statistical test to identify meaningful relationships among the data (i.e., be robust), a sample must be relatively large (e.g., n > 30; Madrigal 1995:9).

This is because the effect of outliers is much greater in small samples, which can exaggerate or obscure trends within the sample (Buescher 1997). In order to control for this, nonparametric statistical analyses were employed in this study because they are less prone to the effects of small sample sizes and abnormally distributed data than parametric

statistical techniques (see Chapter 5).

A total of 21 individuals were identified in the Caledonia collection; bone collagen was extracted from 20 of these individuals, bone bioapatite from 18 and tooth enamel bioapatite from six. The division into subgroups to assess dietary patterns by age, sex, time period and social status further reduced the sample size. For example, poor preservation made the determination of sex for many of the individuals impossible; as a result, only 6 individuals could be assigned a probable sex. In addition, not all of these individuals could be sampled for all types of tissue. One of the probable males was only sampled for tooth enamel bioapatite, one of the probable females was sampled for collagen and bone bioapatite, and another female was sampled for collagen and tooth enamel bioapatite, which further reduced the sample size.

Whittington and Reed (1997:158) explain that small sample size is problematic

"because high variance in small samples allows spurious and contradictory data patterns to arise". A good illustration of this involves a Maya collection analyzed by Reed (1994,

1999). Initially, 25 individuals were sampled from Copan for isotopic analysis, and no dietary differences could be discerned except for a child of nursing age (Reed 1994).

Subsequent sampling of more individuals from the site, however, revealed variation in diet based on sex and social status (Reed 1999). It is therefore possible that if additional individuals were excavated and analyzed from Caledonia, more discernable dietary differences based on age, sex, social status and time period could become apparent.

7.3.2 Representativeness of the Sample

Collections of skeletal remains excavated from archaeological sites are often used to make interpretations about the population that once lived there. Such interpretations, rely on the assumption that the skeletal sample accurately reflects the population from which it was derived. However, several factors can bias skeletal collections including those that cannot be controlled (biological mortality bias, cultural mortality bias, and preservation), and those that can (excavation strategy, recovery factors; Mays 1998:14).

7.3.2.1 Uncontrollable Factors

The 'biological mortality bias' (Saunders and Hoppa 1993:127) refers to the fact that a skeletal sample represents the non-survivors and thus may not accurately reflect the population from which it is derived. This has two implications for this study: (1) the disease processes discussed above may not accurately represent the true disease profile at

Caledonia; and (2) the isotopic results may have been affected by undetectable disease

processes.

To address the first point, individuals must survive a period of stress or disease in

order for it to manifest on their skeleton. Mild episodes of stress during childhood, for

example, will not be expressed as linear enamel hypoplasia in the dentition if they are not

severe enough to disrupt growth. Similarly, anemia must occur during childhood and be

severe and chronic in order for it to produce porotic hyperostosis in the cranium. At

Caledonia, the isotopic values of individuals who expressed these conditions were

compared to those who did not. It was assumed that if the pathological lesions were not visible in the skeleton or dentition, then the individual did not suffer from the causative

disease process. It is, however, possible that individuals who did not express pathology in their skeletons and dentition may have been suffering from disease processes that either

do not manifest physically or were acute and thus did not have time to affect the skeleton

and dentition. Furthermore, it is possible that the skeletal elements on which certain conditions are commonly observed (i.e., porotic hyperostosis on the parietal bones), may not be available for examination due to damage or poor preservation. In summary, the lack of pathological conditions visible in the skeleton does not necessarily mean that the individual was healthy.

Second, studies have demonstrated that different diseases can alter the results of isotopic analysis (Katzenberg and Lovell 1999; White and Armelagos 1997). The repair of fractured bone, for example, involves the deposition of new tissue, which can document short-term dietary change (Katzenberg and Lovell 1999). These studies have logically focused on conditions that are visible in skeletal remains. If pathological 162 conditions that do not manifest skeletally affect the distribution of macronutrients and/or metabolic rates, they are a concern for isotopic interpretation. For example, while investigating the effect of osteomyelitis on isotopic data, Katzenberg and Lovell (1999) found that wasting disease and the consequent recycling of tissue protein can elevate S15N values (Katzenberg and Lovell 1999). Stable nitrogen isotopes are particularly susceptible to physiological factors caused by pathological conditions that are not related to diet

(White and Armelagos 1997). The best way of mitigating the effects of pathological bone on isotopic signatures is simply to sample unaffected bone. Pathological conditions that do not manifest skeletally, however, are much more difficult to detect and control for.

Another concern is known as the 'cultural mortality bias', which takes into account the beliefs of a population and their influence on funerary practices (Saunders and Hoppa 1993:129). As discussed above, the ancient Maya did not inter their dead in cemeteries, and instead buried them in buildings at a site. This reduces the chance of recovering human remains during archaeological investigations. For example, due to time constraints, only three major structures at Caledonia were investigated, usually by excavating a single area (Awe 1985). This is problematic because the potential human remains from other structures at the site were not recovered, and the chances of finding skeletal remains in the structures that were excavated were reduced.

Another reason why the cultural mortality bias is problematic is that the identity of excavated individuals can be difficult to discern. For example, some interments have been related to ancestor veneration (McAnany 1995), and thus the remains likely belonged to an important lineage member at a site. Alternatively, some interments have been identified as sacrificial burials associated with the dedication or rebuilding of a structure at a site, and the individual(s) interred in the burials may instead have been 163

drawn from a lower status segment of the population, as may have been the case for the

child burials at Pacbitun (White et al. 1993:361). Attempts have been made to

differentiate between these types of burials through the analysis of mortuary data. For

example, at Caracol sacrificial victims were identified using burial context and artefacts,

and these individuals were found to have "poorer diets" than other individuals buried in the same residential group (A. Chase et al. 2001:113). The cultural mortality bias, then, is problematic because the burials encountered in excavation may not accurately reflect the

entire population of a site.

A related issue that is perhaps unique to the Maya is the re-entry of burials,

specifically tombs, either for sequential interments of either articulated bodies or bundles,

ritual purposes, or acquiring bones of the ancestors for veneration (see McAnany 1995).

Ritual rites performed at tombs may have little impact on the condition of the bones, but the placement of new individuals in a tomb and the removal of bones will disturb the

remains. At Caledonia, for example, when a more recently deceased individual was added to the burial, the remains of the former individuals may have been pushed to the side to make room for the new body (Healy et al. 1998:269). This not only commingled the remains so that individual identification was difficult, but may also have contributed to the fragmentary nature of the remains from this burial.

Ancient Maya skeletons are notoriously poorly preserved because bone breaks down quickly in the acidic soil and tropical environment of Central America. Additional factors such as microorganisms in the burial environment (Hanson and Buikstra 1987) as well as water availability and temperature (Mays 1998:21) can affect the preservation of

skeletal material at archaeological sites. If the appropriate skeletal elements are not preserved, it may be difficult to estimate demographic aspects such as age and sex. For example, the skeletal remains excavated from Burial #1 at Caledoma "were in extremely poor condition and unrecoverable" (Awe 1985:104). The bones that were recovered from this context were primarily the bones of the hand and feet, from which sex and age estimations are difficult to make.

While poor preservation is certainly a concern, it should not have a negative impact on isotopic studies, as small portions of bone are required for analysis. In this

study, for example, the metatarsals were sampled and yielded adequate collagen for isotopic analysis. However, the rib and cranial samples produced inadequate collagen yields for isotopic analysis. Instead, a second sample of collagen was extracted from the cortical bone of the diaphysis of the left tibia of C2-3-A, C2-3-C and CI-5-A. The high degree of fragmentation of rib and cranial fragments sampled from these individuals would have resulted in increased surface area. This may have facilitated the degradation of these elements to a greater degree than the intact tibiae of these individuals, thus resulting in low collagen yields.

7.3.2.2 Controllable Factors

The excavation strategy and techniques employed can also influence the resulting skeletal assemblage. For example, hand sieving can overlook small bones such those of the hands and feet (Mays 1992), as well as other calcified tissues such as kidney stones

(Morris and Rodgers 1989). The use of screens to further sieve the soil in the burial environment can increase the recovery of small bones as well as evidence of disease preserved in the burial environment. Additionally, in cases where archaeologists do not have direct control over which areas of the site are excavated, or are at the mercy of financial or temporal constraints, the skeletons in the excavation area may not be representative of those at the site as a whole. This can be problematic in areas where differential treatment of the dead was practiced (Mays 1998).

Archaeologists have specific research goals in mind when they are creating their research design. In the past, this has often not been specifically oriented towards the recovery of a representative sample of human skeletal material (among other data), and usually the recovery of a skeleton only occurred by chance. This was the case at

Caledonia, where excavations focused on the creation of structure construction histories and a site chronology (Awe 1985). Furthermore, excavations in the Maya region tended to focus on the larger sites and structures at these sites, which were generally associated with the elite members of society. It is only within the last few decades that researchers have shifted their focus towards the lower strata of society.

7.4 Chapter Seven Summary

This chapter presented detailed interpretations of the isotopic analysis of human bone and tooth samples from Caledonia. The diet of the Caledonia Maya was heavily dependent on maize, although there is evidence for the consumption of terrestrial animals and freshwater molluscs, and four individuals may have consumed marine protein. No statistically significant dietary differences were associated with any of the investigated parameters. While considerable maize consumption during childhood at Caledonia may be associated with the development of anemia-related porotic hyperostosis and dental calculus, the correlation between a maize-rich diet and the expression of linear enamel hypoplasia, dental caries, or dental wear seen at other Maya sites was not identified. It must be emphasized, however, that small sample size limits the conclusions that may be drawn from this data. No significant differences were observed between the Late Preclassic, Early to

Late Classic, and Late Classic samples. It is also interesting that the four individuals who appear to have consumed marine protein (C2-3-B, C2-3-D, C2-4-F1, and C2-4-F6) all date to the Late Classic period. It appears that Burials #1 and #4 represent the tombs of the highest ranking members of society at Caledonia, and Burials #3 and #5 represent dedicatory or commemorative burials in honour of the structures in which they were interred. No dietary differences were observed between the individuals buried in the two tombs, and it appears that they consumed more maize and protein than the individuals interred in the cist and simple burials.

Although several issues, including problems of sample size, as well as controllable and uncontrollable factors related to the representativeness of the sample were acknowledged, an inter-site comparison of isotopic results was conducted to better situate the diet of Caledonia within the Maya region. This comparison found that, isotopically, individuals from Caledonia resembled those from nearby sites such as

Caracol and Pacbitun, as well as those from several sites in the Peten. Unexpectedly, most of the isotopic values from the Caledonia Maya did not closely resemble those at other

Belizean sites, likely due to differences in the locally available subsistence resources. CHAPTER EIGHT CONCLUSION

8.1 Summary and Significance of Research

This study utilized the analysis of stable carbon and nitrogen isotopes in human bones and teeth from Caledonia, Cayo District, Belize, to characterize the diet of the individuals who lived at this site. The biochemical evidence of diet at Caledonia was contextualized using archaeological, ethnohistoric, ethnographic, linguistic, palaeobotanical, and zooarchaeological lines of evidence, and this revealed several general observations that will be reiterated here. First, the Caledonia Maya consumed a . large proportion of maize in their diet, as evidenced by enriched 8 Cbio values from the site. Tooth enamel bioapatite values were even more enriched than bone bioapatite, suggesting that maize was more substantial in the diet of subadults than it was for adults.

Carbon in the protein portion of the diet was somewhat less enriched, suggesting the consumption of primarily maize-fed terrestrial animals, perhaps deer, but also some C3- fed terrestrial animals or freshwater molluscs.

In order to further investigate the type of protein consumed, stable nitrogen isotope analysis and the spacing between stable carbon isotopes in bone collagen and bone bioapatite were assessed. Although the stable nitrogen isotope ratios measured in the

Caledonia Maya are elevated enough to suggest substantial meat consumption, it is unlikely that they consumed large quantities of freshwater fish. Instead, the nitrogen isotope ratios are consistent with the consumption of terrestrial animals (i.e., deer, crab or reptiles), or freshwater molluscs (i.e., snails and mussels). The remains of deer, crab, freshwater mussels and snails have been found at Caledonia (Awe 1985), but because C3- 168 fed terrestrial animals and freshwater molluscs exhibit similar 8 C values, it is not possible to differentiate between these resources using stable isotope analysis at this time.

To differentiate between the consumption of terrestrial and marine animals, the collagen-to-bioapatite spacing values were assessed. The majority of values at Caledonia fell within the omnivorous/carnivorous range and none indicated exclusive herbivory.

I3 The omnivorous A Cbi0-coi values were to be expected considering the variety of food resources utilized by the ancient Maya, and the availability of numerous resources to the residents of Caledonia (see Chapter 4). Surprisingly, four individuals (C2-3-B, C2-3-D,

C2-4-F1, and C2-4-F4) exhibited extremely carnivorous values, which were unexpected and may possibly be the result of marine resource consumption.

No statistically significant patterns related to diet and pathology at Caledonia were discerned. As such, the correlation between a maize-rich diet and the expression of porotic hyperostosis, linear enamel hypoplasia, dental caries, dental calculus, or dental wear seen at other Maya sites was not identified at Caledonia.

Because of small sample sizes, it was difficult to statistically evaluate dietary differences at the site based on age, sex, social status and time period. Some of the individuals sampled consumed more maize as subadults, and the females sampled may have enjoyed a more varied diet and more maize than the males who were analyzed.

While social status was difficult to distinguish at Caledonia, it appears that the tombs represent the two elite lineages at the site; the cist burial may represent some type of ritual or political interment, whereas the simple burial may be dedicatory in nature. The individual interred in the simple burial consumed the least amount of protein and maize, while individuals interred in the tomb consumed the most. Temporally, no significant differences were noted between time periods. Finally, the isotopic data from Caledonia 169

cluster with that of nearby sites such as Pacbitun and Caracol, but the Caledonia values

are more similar to those from sites in Guatemala than to other sites in Belize.

This research is of scholastic importance for several reasons. First, the importance of sampling multiple elements (i.e., carbon and nitrogen) as well as multiple phases of bone and tooth (i.e., bone collagen, bone bioapatite and tooth enamel bioapatite) in order to accurately elucidate individual diets has been demonstrated. Similarly, this study has demonstrated the value of using a multi-evidentiary approach in order to more fully comprehend dietary practices in past societies. Third, this study focuses on the diet of individuals from a minor Maya centre, rather than individuals from the core of large primary centres which have dominated this field of research in the past. As such, it adds to the growing body of isotopic data obtained from other smaller Maya sites, allowing for a comparison with larger sites in order to increase our understanding of regional dietary patterns (Mansell et al. 2006; Metcalfe et al. 2009a; Williams et al. 2009). Fourth, this research has demonstrated that among the ancient Maya diet was not solely determined by locally available resources, but also by political, economic and cultural factors.

Finally, this study illustrates that collagen and bone and tooth enamel bioapatite can be

successfully extracted and analyzed isotopically from skeletal remains that have been curated for some time.

8.2 Limitations of this Study

While it was possible to infer the types of foods consumed by the Caledonia Maya using a variety of sources, the limitations of this study must be reiterated. The sample size from Caledonia was small (n=21), and the division into subgroups based on age, sex,

social status, and time period further reduced the sample size. Small sample sizes make 170

generalizations about the population difficult to make, and can potentially obscure or

exaggerate statistical trends in the data. It is possible that if additional individuals were

excavated from Caledonia and sampled for stable isotope analysis, patterns within the data might become more apparent, as was the case at Copan (Reed 1994, 1999).

Additionally, the representativeness of the sample was affected by both uncontrollable and controllable factors. The biological mortality bias acknowledges that the skeletal sample represents non-survivors who may not accurately represent the population from which they were derived. This is important to recognize because (1) the pathological conditions observed in the skeletons of sampled individuals may not

accurately represent the true disease profile at Caledonia; and (2) the isotopic results may be affected by diseases unobservable in the skeletal remains.

A second issue is the cultural mortality bias, which takes into account the beliefs of a population, which can influence their funerary practices. For example, the ancient

Maya interred both important lineage members as well as dedicatory human burials in the permanent architecture at a site. Furthermore, the re-entry of tombs for sequential burial or ritual purposes can complicate interpretations of the burial context. It is therefore important to carefully review the funerary context in order to interpret the results of isotopic analysis.

While the preservation of the skeletons excavated from Caledonia was good relative to other ancient Maya skeletal collections, many of the elements used to estimate age and determine sex did not survive. Thus, interpretations of the isotopic data in terms of age and sex differences in diet are considered tentative.

Controllable factors that affect the representativeness of the skeletal sample include excavation techniques and research design. Although the recovery techniques employed at Caledoma were not specified (see Awe 1985), the research design focused on establishing a site chronology. Therefore, excavations focused on specific areas of the three major structures at the site, and burials included in other contexts at Caledonia were not encountered.

Finally, factors that affect the analysis of stable carbon and nitrogen isotopes presented potential limitations in this study. For example, environmental conditions such as light intensity and humidity can influence the 813C values in plants, which can be passed on to their consumers. Furthermore, isotopic studies that investigate differences based on age can be influenced by the turnover rate of bone, which is not well established. Similarly, the difference between the isotope ratio in diet and bone collagen

(A Cdiet..Coi) and bioapatite (A Cdiet-bw) is not consistent between studies, which can complicate dietary interpretations. These issues, among others, have been discussed in several sections throughout this thesis (3.2.5, 3.3.4, 3.4 and 7.4), and techniques to mediate their effects have been presented.

8.3 Directions for Future Research

Over the last few decades great strides have been made in terms of better defining stable isotope analysis and recognizing and controlling for associated complicating factors such as diagenesis. This progress is encouraging, and below I outline several areas in which I feel there is great potential for future research. The first area is specific to this particular study, whereas the second and third areas are applicable to more general topics in biochemical investigations of diet in archaeological cultures.

This study has revealed that the stable isotope values measured in individuals interred at Caledonia are more consistent with those from of Guatemala than those of 172

nearby sites in Belize (excluding Caracol and Pacbitun). Depending on which sites are

sampled, it may be possible to identify the migration of people between these sites using

stable oxygen and strontium isotope analysis. A strontium map has been constructed for

the Maya region (Hodell et al. 2004), and strontium as well as oxygen have been used to

trace migrations at several sites (Price et al. 2010; White et al. 2007; Wright 2005;

Wright et al. 2010). Oxygen isotope values are available for the individuals sampled from

Caledonia (see Appendix G), but strontium isotopes ratios were not obtained. It would be

interesting to measure the strontium isotope values of individuals from Caledonia, and

compare both the stable strontium and oxygen isotope rations from Caledonia to other

sites in the Highlands and Northern Lowlands as well as the Southern Lowlands in order

to identify potential migration patterns between these areas. Similarly, the assessment of

the stable oxygen isotope data to investigate migration to Caledonia may help to clarify

the A13Cbio-coi values which suggest that four individuals consumed marine resources.

In ancient Maya studies, chronology is usually based on ceramic sequences, as

was the case at Caledonia (Awe 1985). Such chronologies provide large date ranges,

which are problematic for isotopic investigations where small ranges would be preferred.

Furthermore, in cases where multiple burials are present, as with Burials #1 and #4 at

Caledonia, it would be preferable to know the sequence in which individuals were buried.

Radiocarbon dating of skeletal material to be sampled for stable isotope analysis holds the

potential to further refine site chronology, and better our understanding of dietary change through time. A fellow graduate student at Trent University, Shannen Stronge, has •

incorporated this into her Master's research, and the preliminary results look very promising. 173

A third area of interest involves tracking dietary change over the lifetime of a single individual. In this study, stable carbon isotopes in bioapatite in tooth enamel, deposited during childhood, were compared to those in bone bioapatite, which reflect those deposited within the last several years of an individual's life. In this way, the childhood diet can be contrasted with that from adulthood. A better understanding of the

1 3 diet to tooth enamel spacing (A Cdiet-en) and the turnover rates of different skeletal tissues would greatly improve the potential of such a comparison to track changes in the diet of a single individual.

In order to track changes in diet over the period in which teeth form, Balasse

(2002) analyzed stable carbon and oxygen isotopes in tooth enamel samples from steers

(Bos taurus) raised on a controlled diet. This study revealed that tooth enamel bioapatite in steer was enriched on average by 13.4±0.7%o relative to diet (Balasse 2002), compared to an estimated 12%o proposed for bone bioapatite (Krueger and Sullivan 1984; Lee-

Thorp et al. 1989). At Cuello, van der Merwe and colleagues (2000:31) found that tooth enamel bioapatite was enriched by roughly l%o relative to bone bioapatite, which was attributed to the trophic effect from the pre-weaning diet represented in the early forming 1 ^ teeth. Similar research would help to clarify the average A Cdiet-en value for humans, which would be helpful for dietary studies of archaeological populations.

Finally, because dentine has been found to be more susceptible to diagenetic alteration than tooth enamel (Budd et al. 2000; Koch et al. 1997), it has not often been sampled for isotopic analysis. Recently, however, isotopic values in dentine collagen have been compared to those in dentine and tooth enamel bioapatite to track dietary changes in a single individual over time (Balasse and Tresset 2002; Balasse et al. 2001; Fuller et al.

2003; Koch et al. 1989), particularly to investigate weaning practices in the Maya region 174

(Wright and Schwarcz 1999; Wright et al. 2010). Such sampling can be used to document seasonal variation in diet, which previously was limited to sampling of preserved hair

(Tykot 2002:227). Dentine has also been isotopically sampled in order to investigate migration patterns among the ancient Maya (Wright et al. 2010). It should be cautioned, however, that the addition of secondary dentine in adulthood obscures the signal of childhood diet preserved in primary dentine, which influences dietary interpretation from isotopic analysis of this tissue (Sealy et al. 1993:88). Secondary dentine can be successfully, removed (Wright and Schwarcz 1999), however, and the sampling of primary dentine can provide an additional source of information for dietary studies. REFERENCES CITED

Acsadi G, Nemeskeri J. 1970. History of Human Life Span and Mortality. Budapest: Akademiai Kiado.

Ambrose S H. 1986. Stable carbon and nitrogen isotope analysis of human and animal diet in Africa. Journal of Human Evolution 15:707-731.

Ambrose SH. 1990. Preparation and characterization of bone and tooth collagen for isotopic analysis. Journal of Archaeological Science 17:431 -451.

Ambrose SH. 1991. Effects of diet, climate and physiology on nitrogen isotope abundances in terrestrial foodwebs. Journal of Archaeological Science 18(3):293-317.

Ambrose SH. 1993. Isotopic analysis of paleodiets: Methodological and interpretive considerations. In: Sandford MK, editor. Investigations of Ancient Human Tissue: Chemical Analysis in Archaeology. Langhorne: Gordon and Breach, pp. 59-125.

Ambrose SH. 2000. Controlled diet and climate experiments on nitrogen isotope ratios of rats. In: Ambrose SH, Katzenberg MA, editors. Biogeochemical Approaches to Paleodietary Analysis. New York: Kluwer Academic/Plenum Publishers, pp. 243-259.

Ambrose SH, DeNiro MJ. 1986. The isotope ecology of east African mammals. Oecologia 69:395-406.

Ambrose SH, Norr L. 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: Lambert JB, Grupe G, editors. Prehistoric Human Bone —Archaeology at the Molecular Level. : Springer-Verlag, pp. 1-37.

Ambrose SH, Butler BM, Hanson DB, Hunter-Anderson RL, Krueger HW. 1997. Stable isotopic analysis of human diet in the Marianas Archipelago, Western Pacific. American Journal of Physical Anthropology 104:343-361.

Andrews EWIV, Andrews EW V. 1980. Excavations at , Yucatan, Mexico. Middle American Research Institute Publication 48. New Orleans: Tulane University.

Ankli A, Sticher O, Heinrich M. 1999. Yucatec Maya medicinal plants versus nonmedicinal plants: Indigenous characterization and selection. Human Ecology 27(4):557-580.

Armelagos GJ. 1994. "You are what you eat". In: Sobolik KD, editor. Paleonutrition: The Diet and Health of Prehistoric Americans. Center for Archaeological Investigations, Occasional Paper No. 22. Bethesda, MD: Board of Trustees, Southern Illinois University, pp. 235-244. 176

Asala SA. 2001. Sex determination from the head of the femur of South African whites and blacks. Forensic Science International 117(1): 15-22.

Atran S, Ucan Ek' E. 1999. Classification of useful plants by the northern Peten Maya (Itzaj). In: White CD, editor. Reconstructing Ancient Maya Diet. Salt Lake City: The University of Utah Press, pp. 19-59.

Aufderheide AC, Tieszen LL, Allison MJ, Wallgren J, Rapp G Jr. 1988. Chemical reconstruction of components in complex diets: A pilot study. In: Kennedy BE, LeMoine GM, editors. Diet and Subsistence: Current Archaeological Perspectives. Calgary: University of Calgary Archaeology Association, pp. 301-306.

Awe JJ. 1985. Archaeological Investigations at Caledonia, Cayo District, Belize. Master's Thesis, Peterborough, ON: Trent University.

Balasse M. 2002. Reconstructing dietary and environmental history from enamel isotopic analysis: Time resolution of intra-tooth sequential sampling. International Journal of Osteoarchaeology 12:155-165.

Balasse M, Tresset A. 2002. Early weaning of Neolithic domestic cattle (Bercy, France) revealed by intra-tooth variation in nitrogen isotope ratios. Journal of Archaeological Science 29:853-859.

Balasse M, Bocherens H, Mariotti A, Ambrose SH. 2001. Detection of dietary changes by intra-tooth carbon and nitrogen isotopic analysis: An experimental study of dentine collage of cattle (Bos taurus). Journal of Archaeological Science 28:235-245.

Barns E. 2005. Diseases and Human Evolution. Albuquerque: University of New Mexico Press.

Becker MJ. 1992. Burials as caches; caches as burials: A new interpretation of the meaning of ritual deposits among the Classic Period lowland Maya. In: Danien EC, Sharer RJ, editors. New Theories on the Ancient Maya. University Museum Monograph 77. Philadelphia: University of Pennsylvania, pp. 185-196.

Benson EP. 1974. Gestures and offerings. In: Greene-Robertson M, editor. Primera Mesa Redonda de Part I: A Conference on the Art, Iconography and Dynastic History of Palenque. Pebble Beach, CA: The Robert Louis Stevenson School, pp. 109-120.

Bibby BG. 1961. Cariogenicity of foods. Journal of the American Medical Association 177(5):316-321.

Bocherens H, Drucker D. 2003. Trophic level isotopic enrichment of carbon and nitrogen in bone collagen: Case studies from recent and ancient terrestrial ecosystems. International Journal of Osteoarchaeology 13:46-53. 177

Bocherens H, Fizet M, Mariotti A, Lange-Badre B, Vandermeersch B, Borel JP and Bellon G. 1991. Isotopic biogeochemistry (13C, 15N) of fossil vertebrate collagen: Application to the study of a past food web including Neanderthal man. Journal of Human Evolution 20:481-492.

Bowen WH, Pearson SK. 1993. Effect of milk on cariogenesis. Caries Research 27:461- 466.

Brickley M, Ives R. 2008. The Bioarchaeology of Metabolic Bone Disease. Oxford: Academic Press.

Brooks ST, Suchey JM. 1990. Skeletal age determination based on the os pubis: A comparison of the Ascadi-Nemeskeri and Suchey-Brooks methods. Journal of Human Evolution 5:227-238.

Brothwell DR. 1972. Digging Up Bones: The Excavation, Treatment and Study of Human Skeletal Remains. 2nd ed. London: British Museum of Natural History.

Brown CH. 2010. Prehistoric chronology of the common bean in the New World: The linguistic evidence. In: Staller JE, Carrasco M, editors. Pre-Columbian Foodways: Interdisciplinary Approaches to Food, Culture, and Markets in Ancient Mesoamerica. New York: Springer Science, pp. 273-291.

Brown CT, Witschey WRT. 2008. Blank Map of the Maya Region. The Electronic Atlas . of Ancient Maya Sites. URL: http://mavagis.smv.org/maps of the maya area.htm, accessed: March 12, 2010.

Brown TA, Nelson DE, Cogel JS, Southon JR. 1988. Improved collagen extraction by modified Longin method. Radiocarbon 30:171-177.

Budd P, Montgomery J, Barreiro B, Thomas RG. 2000. Differential diagenesis of strontium in archaeological human dental tissues. Applied Geochemistry 15:687-694.

Buescher PA. 1997. Statistical Primer, No. 12. North Carolina Department of Environment, Health and Natural Resources. Raleigh, N. C: State Centre for Health Statistics. Manuscript in the possession of the author.

Buikstra JE, Ubelaker DH, editors. 1994. Standards for Data Collection from Human Skeletal Remains. Arkansas Archaeological Survey Research Series No. 44. FayetteviUe, AR: Arkansas Archaeological Survey.

Bumsted MP. 1985. Past human behavior from bone chemical analysis - respects and prospects. Journal of Human Evolution 14:539-551.

Calvin M, Benson AA. 1948. The path of carbon in photosynthesis. Science 107:476-480. 178

Carbone VA, Keel BC. 1985. Preservation of plant and animal remains. In: Gilbert RI, Mielke JH, editors. The Analysis of Prehistoric Diets. Orlando: Academic Press, pp. 1-19.

Carlos JP, Gittelsohn AM. 1965. Longitudinal studies of the natural history of caries. II. A life table study of caries incidence in the permanent teeth. Archives of Oral Biology 10:739-751.

Carr HS. 1985 Subsistence and Ceremony: Faunal Utilization in a Late Preclassic Community at Cerros, Belize. Papers of the Peabody Museum of Archaeology and Ethnology No. 77. Cambridge: .

Carr HS. 1986. Faunal Utilization in a Late Community at Cerros, Belize. Ph.D. Dissertation, New Orleans: Tulane University.

Carr HS. 1996. Precolumbian Maya exploitation and management of deer populations. In: Fedick SL, editor. The Managed Mosaic: Ancient Maya Agriculture and Resource Use. Salt Lake City: University of Utah Press, pp. 251-261.

Chase AF. 1992. Elites and the changing organization of Classic Maya society. In: Chase DZ, Chase AF, editors. Mesoamerican Elites: An Archaeological Assessment. Norman: University of Oklahoma Press, pp. 30-49.

Chase AF, Chase DZ. 1987. Investigations at the Classic of Caracol, Belize: 1985-1987. Monograph 3. San Francisco: Pre-Columbian Art Research Institute.

Chase AF, Chase DZ. 1996. The organization and composition of Classic Lowland Maya society: The view from Caracol, Belize. In: Robertson MG, Macri MJ, McHargue J, editors. Eight Palenque Round Table, 1993. San Francisco: Pre-Columbian Art Research Institute, pp. 213-222.

Chase AF, Chase DZ. 2001. The royal court of Caracol, Belize: Its palaces and people. In: Inomata T, Houston SD, editors. Royal Courts of the Ancient Maya, Volume Two. Boulder, CO: Westview Press, pp. 102-137.

Chase AF, Chase DZ, White CD. 2001. El paisaje urbano Maya: La integration de los espacios construdios y la estructura social en Caracol, Belice. In: Ciudad Ruiz A, Ponce de Leon MJI, Martinez Martinez MDC, editors. Reconstruyendo la Ciudad Maya: El Urbanismo en las Sociedades Antiguas. Publicaciones de las S. E. E. M. Num. 6. Madrid: Sociedad Espanola de"Estudios Mayas, pp. 95-122.

Chase DZ. 1994. Human osteology, pathology, and demography as represented in the burials of Caracol, Belize. In: Chase DZ, Chase AF, editors. Studies in the Archaeology of Caracol, Belize. Monograph 7. San Francisco: Pre-Columbian Art Research Institute, 123-138. 179

Chase DZ. 1997. Southern Lowland Maya archaeology and human skeletal remains: Interpretations from Caracol (Belize), Santa Rita Corozal (Belize), and Tayasal (Guatemala). In: Whittington SL, Reed DM, editors. Bones of the Maya: Studies of Ancient Skeletons. Washington, DC: Smithsonian Institution Press, pp. 15-27.

Chase DZ, Chase AF, editors. 1992. Mesoamerican Elites: An Archaeological Assessment. Norman: University of Oklahoma Press.

Cheetham D. 2010. Corn, colanders, and cooking: Early maize processing in the Maya Lowlands and its implications. In: Staller JE, Carrasco M, editors. Pre-Columbian Foodways: Interdisciplinary Approaches to Food, Culture, and Markets in Ancient Mesoamerica. New York: Springer Science, pp. 345-368.

Chisholm BS, Nelson DE, Schwarcz HP! 1982. Stable-carbon isotope ratios as a measure of marine versus terrestrial protein in ancient diets. Science 216(4550):1131-1132.

Chisholm BS, Nelson DE, Hobson KA, Schwarcz HP, Knyf M. 1983. Carbon isotope measurement techniques for bone collagen: Notes for the archaeologist. Journal of Archaeological Science 10:355-360.

Christenson A J. 2010. Maize was their flesh: Ritual feasting in the Maya Highlands. In: Staller JE, Carrasco M, editors. Pre-Columbian Foodways: Interdisciplinary Approaches to Food, culture, and Markets in Ancient Mesoamerica. New York: Springer Science, pp. 577-600.

Clementz MT, Fox-Dobbs K, Wheatley PV, Koch PL, Doak DF. 2009. Revisiting old bones: coupled carbon isotope analysis of bioapatite and collagen as an ecological and palaeoecological tool. Geological Journal 44:605-620.

Coe SD. 1994. America's First Cuisines. Austin: University of Texas Press.

Coe WR. 1959. Piedras Negras Archaeology: Artifacts, Caches, and Burials. Philadelphia: The University Museum, University of Pennsylvania.

Cohen MN, Armelagos GJ, editors. 1984. Paleopathology at the Origins of Agriculture. New York: Academic Press.

Coplen TB. 1994. Reporting of stable hydrogen, carbon, and oxygen isotopic abundances. Pure and Applied Chemistry 66(2):273-276.

Costa RL. 1980. Incidence of caries and abscesses in archaeological Eskimo skeletal sample from Point Hope and Kodiak Island, Alaska. American Journal of Physical Anthropology 52:501-514.

Coyston S. 1995. An Application of Carbon Isotope Analysis of Bone Apatite to the Study of Maya Diets and Subsistence at Pacbitun and Lamanai, Belize. Master's Thesis, Peterborough, ON: Trent University. 180

Coyston S, White CD, Schwarcz HP. 1999. Dietary carbonate analysis of bone and enamel for two sites in Belize. In: White CD, editor. Reconstructing Ancient Maya Diet. Salt Lake City: University of Utah Press, pp. 221-243.

Crane CJ. 1996. Archaeobotanical and palynological research at a Late Preclassic Maya community, Cerros, Belize. In: Fedick SL, editor. The Managed Mosaic: Ancient Maya Agriculture and Resource Use. Salt Lake City: University of Utah Press, pp. 262-277.

Crane CJ, Carr HS. 1994. The integration and quantification of economic data from a Late Preclassic Maya community in Belize. In: Sobolik KD, editor. Paleonutrition: The Diet and Health of Prehistoric Americans. Bethesda, MD: Board of Trustees, Southern Illinois University, pp. 66-79.

Crist RE. 1939. The pulque industry. Economic Geography 15(2): 189-194.

Cucina A, Tiesler V. 2003. Dental caries and antemortem tooth loss in the northern Peten area, Mexico. A biocultural perspective on social status differences among the Classic Maya. American Journal of Physical Anthropology 122(1):1-10.

Cucina A, Tiesler Bios V, Sierra Sosa T. 2003. Sex differences in oral pathologies at the Late Classic Maya site of Xcambo, Yucatan. Dental Anthropology 16(2):45-51.

Delmas PD. 1993. Biochemical markers of bone turnover. Journal of Bone and Mineral Research 8:S549-S555.

Delwiche CC, Steyn PL. 1970. Nitrogen isotope fractionation in soils and microbial reactions. Environmental Science and Technology 4:929-935.

Dembo A, Imbelloni J. 1938. Deformaciones intencionales del cuerpo humano de character etnico. Buneos Aires: Biblioteca Humanior.

DeNiro MJ. 1985. Postmortem preservation and alteration of in vivo bone collagen isotope ratios in relation to palaeodietary reconstruction. Nature 317(31):806-809.

DeNiro MJ. 1987. Stable isotopy and archaeology. American Scientist 75:182-191.

DeNiro MJ, Epstein S. 1978. Influence of diet on the distribution of carbon isotopes in animals. Geochimica et Cosmochimica Acta 42:495-506.

DeNiro MJ, Epstein S. 1981. Influence of diet on the distribution of nitrogen isotopes in animals. Geochimica et Cosmochimica Acta 45:341-351.

DeNiro MJ, Schoeninger MJ. 1983. Stable carbon and nitrogen isotope ratios of bone collagen: Variations within individuals, between sexes, and within populations raised on monotonous diets. Journal of Archaeological Science 10:199-203. 181

DeNiro MJ, Weiner S. 1988. Chemical, enzymatic and spectroscopic characterization of "collagen" and other organic fractions from prehistoric bones. Geochimica et Cosmochimica Acta 52:2197-2206.

Diaz del Castillo B. 1927. The True History of the Conquest of Mexico. Keatinge M, translator. New York: McBride.

Dittrick J, Suchey JM. 1986. Sex determination of prehistoric central California skeletal remains using discriminant analysis of the femur and humerus. American Journal of Physical Anthropology 70(l):3-9.

Dunning NP. 1996. A reexaminaiton of regional variability in the Pre-Hispanic agricultural landscape. In: Fedick SL, editor. The Managed Mosaic: Ancient Maya Agriculture and Resource Use. Salt Lake City: University of Utah Press, pp. 53-68.

El-Kammar A, Hancock RGV, Allen RO. 1989. Human bones as archaeological samples: Changes due to contamination and diagenesis. In: Allen RO, editor. Archaeological Chemistry IV. Advances in Chemistry Series 220. Washington, DC: American Chemical Society.

El-Najjar MY. 1977. Porotic hyperostosis in North America: A theory. In: Cockburn E, editor. Porotic Hyperostosis: An Enquiry, Monograph No. 2. Detroit, MI: Paleopathology Association, pp. 9-10.

Elliott JC. 1994. Structure and Chemistry of the Apatites and Other Calcium Orthophosphates. Studies in Inorganic Chemistry 18. Amsterdam: Elsevier.

Emery KF. 2003. The noble beast: Status and differential access to animals in the Maya world. World Archaeology 34(3):498-515.

Emery KF, editor. 2004. Maya Zooarchaeology: New Directions in Method and Theory. Costen Institute of Archaeology, Monograph 51. Los Angeles: Cotsen Institute of Archaeology.

Emery KF. 2010. Dietary, Environmental, and Societal Implications of Ancient Maya Animal Use in the Petexbatun: A Zooarchaeological Perspective on the Collapse. Vanderbilt Institute of Mesoamerican Archaeology Vol. 5. Nashville: Press.

Emery KF, Wright LE, Schwarcz H. 2000. Isotopic analysis of ancient deer bone: Biotic stability in collapse period Maya land-use. Journal of Archaeological Science 27:537- 550.

Evans TD. 1973. A preliminary evaluation of tooth tartar among the Preconquest Maya of the Tayasal Area, El Peten, Guatemala. American Antiquity 38:489-493. 182

Farriss NM. 1984. Maya society Under Colonial Rule: The Collective Enterprise of Survival. Princeton; Princeton University Press.

Fash WL, Fash BW, Davis-Salazar KL. 2004. Setting the stage: Origins of the hieroglyphic stairway plaza on the great period ending. In: Bell EE, Canuto MA, and Sharer RJ, editors. Understanding Early Classic Copan. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology, pp. 65-83.

Faure G. 1986. Principles of Isotope Geology, Second Edition. New York: John Wiley and Sons.

Fogel ML, Tuross N, Owsley DW. 1989. Nitrogen isotope tracers of human lactation in modern and archaeological populations. Carnegie Institution, annual report of the director, Geophysical Laboratory, pp. 111-117.

Franco JL. 1969. Snares and Traps in Codex Madrid. Notes on Middle American Archaeology and Ethnology 5(121). Washington, DC: Carnegie Institution of Washington. Reprinted in New York: AMS Press.

Friedli DA, Escobedo HL, Guenter SP. 2007. A crossroads of conquerors: Waka' and Gordon Willey's "rehearsal for the collapse" hypothesis. In: Sabloff JA, Fash WL, editors. Gordon R. Willey and American Archaeology: Contemporary Perspectives. Norman: University of Oklahoma Press, pp. 187-208.

Friedli H, Lotscher H, Oescheger H, Siegenthaler U, Stauffer B. 1986. Ice core record of 13 12 the C/ C ratio of atmospheric C02 in the past two centuries. Nature 324:237-238.

Fuller BT, Richards MP, Mays SA. 2003. Stable carbon and nitrogen isotope variations in tooth dentine serial sections from Wharram Percy. Journal of Archaeological Science 30:1673-1684.

Fuller BT, Fuller JL, Sage NE, Harris DA, O'Connell TC, Hedges REM. 2004. Nitrogen balance and S15N: why you're not what you eat during pregnancy. Rapid Communications in Mass Spectrometry 18:2889-2896.

Fuller BT, Fuller JL, Sage NE, Harris DA, O'Connell TC, Hedges REM. 2005. Nitrogen balance and S15N: Why you're not what you eat during nutritional stress. Rapid Communications in Mass Spectrmetery 19:2497-2506.

Fuller BT, MoUeson, Harris DA, Gilmour LT, Hedges REM. 2006. Isotopic evidence for breastfeeding and possible adult dietary differences from Late/Sub-Roman Britain. American Journal of Physical Anthropology 129:45-54.

Garvie-Lok S. 2001. Loaves and Fishes: A Stable Isotope Reconstruction of Diet in Medieval . Ph.D. Dissertation, Calgary, AB: University of Calgary. 183

Garvie-Lok SJ, Varney TL, Katzenberg MA. 2004. Preparation of bone carbonate for stable isotope analysis: The effects of treatment time and acid . Journal of Archaeological Science 31:763-776.

Gerry JP. 1993. Diet and Status Among the Classic Maya: An Isotopic Perspective. Ph.D. Dissertation, Cambridge, USA: Harvard University.

Gerry JP. 1997. Bone isotope ratios and their bearing on elite privilege among the Classic Maya. Geoarchaeology: An International Journal 12(l):41-69.

Gerry JP, Krueger HW. 1997. Regional Diversity in Classic Maya Diets. In: Whittington S, Reed D, editors. Bones of the Ancestors: Recent Studies of Ancient Maya Skeletons. Washington, DC: Smithsonian Institution Press, pp. 196-207.

Gillespie SD. 2001. Personhood, agency, and mortuary ritual: A case study for the ancient Maya. Journal of Anthropological Archaeology 20:73-112.

Glassman DM, Garber JF. 1999. Land use, diet, and their effects on the biology of the prehistoric Maya of Northern Ambergris Cay, Belize. In: White CD, editor. Reconstructing Ancient Maya Diet. Salt Lake City: The University of Utah Press, pp. 119-132.

Glimcher MJ, Bonar LC, Grynpas MD, Landis WJ, Roufosse AH. 1981. Recent studies of bone mineral: Is the amorphous calcium phosphate theory valid? Journal of Crystal Growth 53:100-119.

Gordon CG, Buikstra JE. 1981. Soil pH, bone preservation and sampling bias at mortuary sites. American Antiquity 46:566-571.

Grupe G, Balzer A, Turban-Just S. 2000. Modeling protein diagenesis in ancient bone: Towards a validation of stable isotope data. In: Ambrose SH, Katzenberg MA, editors. Biogeochemical Approaches to Paleodietary Analysis. New York: Kluwer Academic/Plenum Publishers, pp. 173-187.

Gumerman GIV. 1997. Food and complex societies. Journal of Archaeological Method and Theory 4(2): 105-139.

Hageman JB, Goldstein DJ. 2009. An integrated assessment of archaeobotanical recovery methods in the Neotropical rainforest of northern Belize: Flotation and dry screening. Journal of Archaeological Science 36:2841-2852.

Hamblin NL. 1984. Animal Use by the Cozumel Maya. Tucson: University of Arizona Press.

Hamblin NL. 1985. The role of marine resources in the Maya economy: A case study from Cozumel, Mexico. In: Pohl M, editor. Prehistoric Lowland Maya Environment and Subsistence Economy. Cambridge: Harvard University, pp. 159-173. 184

Hamblin NL, Rea AM. 1985. Isla Cozumel archaeological avifauna. In: Pohl M, editor. Prehistoric Lowland Maya Environment and Subsistence Economy. Papers of the Peabody Museum of Archaeology and Ethnology, no. 77. Cambridge: Harvard University, pp. 175-192.

Hammond N. 1978. The myth of the milpa: Agricultural expansion in the Maya Lowlands. In: Harrison PD, Turner BL II, editors. Pre-Hispanic Maya Agriculture. Albuquerque: University of New Mexico Press, pp. 23-34.

Hammond N. 2001. Los orienes de la culura maya y la formation de comunidades rurales. In: Grube N, Eggebrecht E, Seidel M, editors. Los Mayas: Una Civilizacion Milenaria. Colonia: Konemann, pp. 35-47.

Hammond N, Miksicek CH. 1981. Ecology and economy of a Formative Maya site at Cuello, Belize. Journal of Field Archaeology 8:259-269.

Hanson DB, Buikstra JE. 1987. Histomorphological alteration in buried human bone from the Lower Illinois Valley: Implications for palaeodietary research. Journal of Archaeological Science 14:549-563.

Hare PE. 1980. Organic geochemistry of bone and its relation to the survival of bone in the natural environment. In: Behrensmeyer AK, Hill AP, editors. Fossils in the Making. Chicago: University of Chicago Press, pp. 197-207.

Harris EF, McKee JH. 1990. Tooth mineralization standards for Blacks and Whites from the middle southern United States. Journal of Forensic Sciences 34:859-872.

Harrison PD. 1999. The Lords of Tikal: Rulers of an Ancient Maya City. London: Thames and Hudson.

Hassen AA. 1975. Geochemical and Mineralogical Studies on Bone Material and their Implications for Radiocarbon Dating. PhD. Dissertation, Dallas, TX: Southern Methodist University.

Hatch MD, Slack CR. 1966. Photosynthesis by sugarcane leaves: A new carboxylation reaction and the pathway of sugar formation. The Biochemical Journal 101:103-111.

Hatch MD, Slack CR, Johnson HS. 1967. Further studies on a new pathway of photosynthetic carbon dioxide fixation in sugarcane, and its occurrence in other species. The Biochemical Journal 102:411-422.

Hawkes JG. 1998. Back to Vavilov: Why were plants domesticated in some areas and not in others? In: Damania AB, Valkoun J, Willcox G, Qualset CO, editors. The Harlan Symposium. Aleppo, Syria. Available at: http://www2.bioversityinternational.org/publications/Web version/47/ 185

Healy PF. 1990. Excavations at Pacbitun, Belize: Preliminary report on the 1986 and 1987 investigations. Journal of Field Archaeology 17(3):247-262.

Healy PF, Awe JJ, Helmuth H. 1998. An ancient Maya multiple burial at Caledonia, Cayo District, Belize. Journal of Field Archaeology 25(3):261-274.

Healy PF, Emery K, Wright LE. 1990. Ancient and modem Maya exploitation of the Jute snail (Pachychilus). Latin American Antiquity 1(2):170-183. Healy PF, Lambert JDH, Arnason JT, Hebda RJ. 1983. Caracol, Belize: Evidence of ancient Maya agricultural terraces. Journal of Field Archaeology 10(4):397-410.

Healy PF, van Waarden C, Anderson TJ. 1980. Nueva evidencia de antiguas terrazas Mayas en Belice. American Indigena 40:773-796.

Heaton THE, Vogel JC, von la Chevalleriie G, CoUett G. 1986. Climatic influence on the isotopic composition of bone nitrogen. Nature 322:822-823.

Hedges REM. 2003. On bone Collagen - apatite-carbonate isotopic relationships. International Journal of Osteoarchaeology 13:66-79.

Hedges REM, Clement JG, Thomas CDL, O'Connell TC. 2007. Collagen turnover in the adult femoral mid-shaft: Modeled from anthropogenic radiocarbon tracer measurements. American Journal of Physical Anthropology 133:808-816.

HellmuthN. 1977. Cholti-Lacandon Chiapas and Peten Itza agriculture, settlement pattern, and population. In: Hammond N, editor. Social Process in Maya Prehistory: Studies in Memory of Sir Eric Thompson. London: Academic Press, pp. 421-448.

Helmuth H. 1985. Report on the osteology of human remains from Burials 1, 3, 4, and 5 from Caledonia. In: Awe JJ, Archaeological Investigations at Caledonia, Cayo District, Belize. Master's Thesis, Peterborough, ON: Trent University.

Henderson HH. 1998. The Organization of Staple Crop Production in Middle Formative, Late Formative, and Classic Period Farming Households at K'axob, Belize. Ph.D. Dissertation, Pittsburgh, PA: University of Pittsburgh.

Henderson H. 2003. The organization of staple crop production at K'axob, Belize. Latin American Antiquity 14(4):469-496.

Herbert V, Zalusky R. 1962. Interrelations of vitamin B12 and folic acid metabolism: Folic acid clearance studies. Journal of Clinical Investigation 41(6):1263-1276.

HiUson SW. 1979. Diet and dental disease. World Archaeology 2:147-162.

HiUson SW. 1996. Dental Anthropology. Cambridge: Cambridge University Press. 186

Hodell DA, Quinn RL, Brenner M, Kamenov G. 2004. Spatial variation of strontium isotopes ( Sr/ Sr) in the Maya region: A tool for tracking ancient human migration. Journal of Archaeological Science 31:585-601.

Hoefs J. 2004. Stable Isotope Geochemistry, 5th ed. Berlin: Springer-Verlag.

Holland TD, O'Brien MJ. 1997. Parasites, porotic hyperostosis, and the implications of changing perspectives. American Antiquity 62(2): 183-193.

Howland MR, Corr LT, Young SMM, Jones V, Jim S, van der Merwe NJ, Mitchell AD, Evershed RP. 2003. Expression of the dietary isotope signal in the compound-specific 813C values of pig bone lipids and amino acids. International Journal of Osteoarchaeology 13:54-65.

Hus-Ashmore R, Goodman AH, Armelagos GJ. 1982. Nutritional inference from paleopathology. Advances in Archaeological Method and Theory 5:395-474.

Imbelloni J. 1937. Deformaciones Intencionales del Craneo en Sud America. Helmintologia 6:330-406.

Jackson PC, Meinzer FC, Goldstein G, Holbrook NM, Caveller J, Rada F. 1993. Environmental and physiological influences on carbon isotope composition of gap and understory plants in a lowland tropical forest. In: Ehleringer JR, Hall AE, Farquhar GD, editors. Stable Isotopes and Plant Carbon-Water Relations. San Diego: Academic Press, pp. 131-140.

Jim S, Ambrose SH, Evershed RP. 2004. Stable carbon isotopic evidence for differences in the dietary origin of bone cholesterol, collagen and apatite: Implications for their use in palaeodietary reconstruction. Geochimica et Cosmochimica Acta 68(l):61-72.

Johnson MS, Chaffey DR. 1973. An Inventory of the Chiquibul Forest Reserve, Belize. Land Resource Study 14. Tolworth Tower, Surbiton, Surrey: Land Resources Division.

Jorkov MLS, Heinemeier J, Lynnerup N. 2007. Evaluating bone collagen extraction methods for stable isotope analysis in dietary studies. Journal of Archaeological Science 34:1824-1829.

Katz SH, Hediger ML, Valleroy LA. 1974. Traditional maize processing techniques in the New World. Science 184(4138):765-773.

Katzenberg MA. 1988. Stable isotope analysis of animal bone and the reconstruction of human palaeodiet. In: Kennedy BV, LeMone GM, editors. Diet and Subsistence: Current Anthropological Perspectives. Calgary: The University of Calgary Archaeological Association, pp. 307-314. 187

Katzenberg MA. 2008. Stable isotope analysis: A tool for studying past diet, demography, and life history. In: Katzenberg MA, Saunders SR, editors. Biological Anthropology of the Human Skeleton, 2nd ed. Hoboken, N.J.: John Wiley and Sons, pp. 413-441.

Katzenberg MA, Lovell NC. 1999. Stable isotope variation in pathological bone. International Journal of Osteoarchaeology 9:316-324.

Katzenberg MA, Saunders SR, editors. 2008. Biological Anthropology of the Human Skeleton, 2nd ed. Hoboken, New Jersey: John Wiley & Sons, Inc.

Keegan WF, DeNiro MJ. 1988. Stable carbon- and nitrogen-isotope ratios of bone collagen used to study coral-reef and terrestrial components of prehistoric Bahamian diet. American Antiquity 53(2):320-336.

Keeling CD, Mook WG, Tans PP. 1979. Recent trends in the 13C/12C ratio of atmospheric carbon dioxide. Nature 277:121-123

Keenleyside A, Schwarcz HP, Panayotova K. 2011. Oxygen isotopic evidence of residence and migration in a Greek colonial population on the Black Sea. Journal of Archaeological Science (in press).

Kellner CM, Schoeninger MJ. 2007. A simple carbon isotope model for reconstructing prehistoric human diet. American Journal of Physical Anthropology 133:1112-1127.

Koch PL, Fischer DC, Dettman D. 1989. Oxygen isotope variation in the tusks of extinct proboscideans: A measure of season of death and seasonality. Geology 17:515-519.

Koch PL, Tuross N, Fogel ML. 1997. The effects of sample treatment and diagenesis on the isotopic integrity of carbonate in biogenic hydroxylapatite. Journal of Archaeological Science 24:411-429.

Kohn MJ, Schoeninger MJ, Barker WW. 1999. Altered states: Effects of diagenesis on fossil tooth chemistry. Geochimica et Cosmochimica Acta 63(18):2737-2747.

Kreshover SJ. 1960. Metabolic disturbances in tooth formation. Annals of the New York Academy of Sciences 85:161-167.

Kreshover SJ. 1960. Metabolic disturbances in tooth formation. Annals of the New York A cademy of Sciences 85:161-167.

Krueger HW. 1991. Exchange of carbon with biological apatite. Journal of Archaeological Science 18:355-361.

Krueger HW, Sullivan CH. 1984. Models for carbon isotope fractionation between diet and bone. In: Turnlund JE, Johnson PE, editors. Stable Isotopes in Nutrition. Symposium Series 258. Washington, DC: American Chemical Society, pp. 205-222. Lambert JDH, Arnason JT. 1982. Ramon and Maya ruins: An ecological not an economical relation. Science 216:298-299.

Lambert JB, Vlasak Simpson S, Szpunar CB, Buikstra JE. 1985. Bone diagenesis and dietary analysis. Journal of Human Evolution 14:477-482.

Lambert JB, Xue L, Buikstra JE. 1989. Physical removal of contaminative inorganic material from buried human bone. Journal of Archaeological Science 16:427-436.

Larsen CS. 1987. Bioarchaeological interpretations of subsistence economy and behavior from human skeletal remains. Advances in Archaeological Method and Theory 10:339- 445.

LeCount LJ. 2001. Like water for chocolate: Feasting and political ritual among the Late Classic Maya at , Belize. American Anthropologist 103(4):935-953.

Lee-Thorp JA. 2000. Preservation of biogenic carbon isotopic signals in Plio Pleistocene bone and tooth mineral. In: Ambrose SH, Katzenberg MA, editors. Biogeochemical Approaches to Paleodietary Analysis. New York: Kluwer Academic/Plenum Publishers, pp. 89-115.

Lee-Thorp J, Sponheimer M. 2003. Three case studies used to reassess the reliability of fossil bone and enamel isotope signals for paleodietary studies. Journal of Anthropological Archaeology 22:208-216.

Lee-Thorp JA, van der Merwe NJ. 1987. Carbon isotope analysis of fossil bone apatite. South African Journal of Science 83:712-715.

Lee-Thorp JA, van der Merwe NJ. 1991. Aspects of the chemistry of modem and fossil biological apatites. Journal of Archaeological Science 83:712-715.

Lee-Thorp JA, Sealy JC, van der Merwe NJ. 1989. Stable carbon isotope ratio differences between bone collagen and bone apatite, and their relationship to diet. Journal of Archaeological Science 16:585-599.

Lentz DL. 1991. Maya diets of the rich and poor: Paleoethnobotanical evidence from Copan. Latin American Antiquity 2:269-287.

Lentz DL. 1999. Plant resources of the ancient Maya: The paleoethnobotanical evidence. In: White CD, editor. Reconstructing Ancient Maya Diet. Salt Lake City: University of Utah Press, pp. 3-18.

Lentz DL, Beaudry-Corbett M, Reyna de Aguilar ML, Kaplan L. 1996. Foodstuffs, forests, fields, and shelter: A paleoethnobotanical analyses of vessel contents from the Ceren Site, El Salvador. Latin American Antiquity 7:247-262. Libby WF, Berger R, Mead JF, Alexander GV, Ross JF. 1964. Replacement rates for human tissue from atmospheric radiocarbon. Science 146(3648):1170-1172.

Lieverse AR. 1999. Diet and the aetiology of dental calculus. International Journal of Osteoarchaeology 9:219-232.

Longin R. 1971. New method of collagen extraction for radiocarbon dating. Nature 230:241-242.

Lovejoy CO, Meindl RS, Pryzbeck TR, Mensforth RP. 1985. Chronological metamorphosis of the auricular surface of the ilium: A new method for the determination of age at death. American Journal of Physical Anthropology 68:15-28.

Lukacs JR. 2008. Fertility and agriculture accentuate sex differences in dental caries rates. Current Anthropology 49(5):901-914.

Lundell CL. 1933. The agriculture of the Maya. Southwest Review 19:65-77.

Lundell CL. 1938. Plants probably utilized by the Old Empire Maya of Peten and adjacent lowlands. Papers of the Michigan Academy of Sciences, Arts, and Letters 24:37- 56.

Macko SA, Lee WY, Parker PL. 1982. Nitrogen and carbon isotope fractionation by two species of marine amphipods: Laboratory and field studies. Journal of Experimental and Ecology 63:145-149.

Madrigal L. 1995. Statistics for Anthropology. Cambridge: Cambridge University Press.

Magennis AL. 1999. Dietary change at the lowland Maya site of Kichpanha, Belize. In: White CD, editor. Reconstructing Ancient Maya Diet. Salt Lake City: University of Utah Press, pp. 133-150.

Mann RW, Hunt DR. 2005. Photographic Regional Atlas of Bone Disease: A Guide to Pathological and Normal Variation in the Human Skeleton. Springfield: Charles C. Thomas.

Mansell EB, Tykot RH, Freidel DA, Dahlin BH, Ardren T. 2006. Early to Terminal Classic Maya diet in the Northern Lowlands of the Yucatan (Mexico). In: Staller J, Tykot R, Benz B, editors. Histories of Maize: Multidisciplinary Approaches to the Prehistory, Linguistics, Biogeography, Domestication, and Evolution of Maize. Amsterdam: Elsevier, pp. 173-485.

Marcus J. 1982. The plant world of the sixteenth- and seventeenth-century lowland Maya. In: Flannery KV, editor. Maya Subsistence: Studies in Honor of Dennis Puleston. New York: Academic Press, pp. 239-273. Marino BD, McElroy MB. 1991. Isotopic composition of atmosphenc O2 inferred from carbon in C4 plant cellulose. Nature 349:127-131.

Mariotti A. 1984. Natural 15N abundance measurements and atmospheric nitrogen standard calibration. Nature 311:251-252.

Martin DL, Goodman AH, Armelagos GJ. 1985. Skeletal pathologies as indicators of quality and quantity of diet. In: Gilbert RI Jr, Mielke JH, editors. The Analysis of Prehistoric Diets. Orlando: Academic Press, pp. 227-279.

Masson MA, Lope.CP. 2008. Animal use at the Postclassic Maya center of . Quaternary International 191:170-183.

Mays S. 1992. Taphonomic factors in a human skeletal assemblage. Circaea 9:54-58.

Mays S. 1998. The Archaeology of Human Bones. London: Routledge.

McAnany PA. 1995. Living with the Ancestors: Kinship and Kingship in Ancient Maya Society. Austin: University of Texas Press.

McGee RJ. 2002. Watching Lacandon Maya Lives. Boston: Allyn and Bacon.

McKillop H. 1995. , salt production, and coastal Maya trade at Stingray Lagoon, Belize. Latin American Antiquity 6(3):214-228.

McKillop H. 2006. The Ancient Maya: New Perspectives. New York: W. W. Norton and Company.

McKillop H. 2010. Ancient Maya canoe navigation and its implications for Classic to Postclassic Maya economy and sea trade: A view from the south coast of Belize. Journal of Caribbean Archaeology, Special Publication 3: 93-105.

McKillop H, Sabloff J A. 2005. Finds in Belize document Late Classic Maya salt making and canoe transport. Proceedings of the National Academy of Sciences of the United States of America 102(15):5630-5634.

Metcalfe JZ, White CD, Longstaffe FJ, Wrobel G, Cook DC, Pybum KA. 2009a. Isotopic evidence for diet at Chau Hixx, Belize: Testing regional models of hierarchy and heterarchy. Latin American Antiquity 20(l):15-36.

Metcalfe JZ, Longstaffe FJ, White CD. 2009b. Method-dependent variations in stable isotope results for structural carbonate in bone bioapatite. Journal of Archaeological Science 36:110-121.

Meindl RS, Lovejoy CO. 1989. Age changes in the pelvis: Implications for paleodemography. In: Iscan MY, editor. Age Markers in the Human Skeleton. Springfield, IL: Charles C. Thomas, pp. 137-168. 191

Miksicek CH. 1983. Macrofloral remains of the Pulltrouser area: Settlements and fields. In: Turner BL, Harrison PD, editors. : Ancient Maya Habitat, Agriculture, and Settlement. Austin: University of Texas Press, pp. 94-104.

Miksicek CH. 1991. The natural and cultural landscape of Preclassic Cuello. In: Hammond N, editor. Cuello: An Early Maya Community in Belize. Cambridge: Cambridge University Press, pp. 70-84.

Miksicek CH, Bird RM, Pickersgill B, Donaghey S, Cartwright J, Hammond N. 1981. Preclassic lowland maize from Cuello, Belize. Nature 289:56-59.

Moholy-Nagy H. 2003. The hiatus at Tikal, Guatemala. Ancient Mesoamerica 14:77-83.

Moore WJ, Corbett ME. 1971. The distribution of dental caries in ancient Biritsh populations 1: Anglo-Saxon period. Caries Research 5:151 -168.

Moorrees CFA, Fanning EA, Hunt EE. 1963 a. Formation and resorption of three deciduous teeth in children. American Journal of Physical Anthropology 21:205-213.

Moorrees CFA, Fanning EA, Hunt EE. 1963b. Age formation by stages for ten permanent teeth. Journal of Dental Research 42:1490-1502.

Morehart CT, Lentz DL, Prufer KM. 2005. Wood of the gods: The ritual use of pine (Pinus spp.) by the ancient Lowland Maya. Latin American Antiquity 16(3):255-274.

Morley S. 1946. The Ancient Maya, 1st ed. Stanford: Stanford University Press.

Morris AG, Rodgers AL. 1989. A probable case of prehistoric kidney stone disease from the Northern Cape Province, South Africa. American Journal of Physical Anthropology 79:521-527.

Nations, JD. 1979. Snail shells and maize preparation: A Lacandon Maya analogy. American Antiquity 44(3):568-571:

Nations JD, RB Nigh. 1980. The evolutionary potential of Lacandon Maya sustained yield tropical forest agriculture. Journal of Anthropological Research 36:1-30.

Navia JM. 1994. Carbohydrates and dental health. American Journal of Clinical Nutrition 59:719s-727s.

Nelson BK, DeNiro MJ, Schoeninger MJ, De Paolo DJ. 1986. Effects of diagenesis on strontium, carbon, nitrogen and oxygen concentration and isotopic composition of bone. Geochimica et Cosmochimica Acta 50:1941-1949.

Nielson-Marsh CM, Hedges REM. 2000a. Patterns of diagenesis in bone I: The effects of site environments. Journal of Archaeological Science 27:1139-1150. 192

Nielson-Marsh CM, Hedges REM. 2000b. Patterns of diagenesis in bone II: Effects of acetic acid treatment and the removal of diagenetic C02_3. Journal of Archaeological Science 27:1151-1159.

Nitsch EK, Humphrey LT, Hedges REM. 2010. The effect of parity status on 815N: looking for the "pregnancy effect" in 18th and 19th century London. Journal of Archaeological Science 37:3191-3199.

Norr L. 1991. Nutritional Consequences of Prehistoric Subsistence Strategies in Lower Central America.. Ph.D. Dissertation, Urbana-Campagin: University of Illinois

Norr L. 1995. Interpreting dietary maize from bone stable isotopes in the American tropics: The state of the art. In: Stahl PW, editor. Archaeology in the Lowland American Tropics: Current Analytical Methods and Recent Applications. Cambridge: Cambridge University Press, pp. 198-223.

O'Leary M. 1988. Carbon isotopes in photosynthesis. BioScience 38(5):328-336.

Olsen SJ. 1972. Animal remains from Altar de Sacrificios. In: Willey GR, The Artifacts of Altar de Sacrificos. Papers of the Peabody Museum of Archaeology and Enthology, no. 64. Cambridge: Harvard University, pp. 243-246.

Olsen SJ. 1978. Vertebrate faunal remains. In: Willey GR, editor. Excavations at Seibal, Department of Peten, Guatemala. Memoirs of the Peabody Museum of Archaeology and Ethnology 14(1):172-176. Cambridge: Harvard University.

Ortner DJ, editor. 2003. Identification of Pathological Conditions in Human Skeletal Remains, 2nd ed. San Diego: Academic Press.

Ortner DJ, Aufderheide AC, editors. 1991. Human Paleopathology: Current Syntheses and Future Options. Washington, D. C: Smithsonian Institution Press. Palkovich AM. 1987. Endemic disease patterns in paleopathology: Porotic hyperostosis. American Journal of Physical Anthropology. 74:527-537.

Pallant J. 2001. SPSS Survival Manual: A Step by Step Guide to Data Analysis Using SPSS for Windows (Versions 10 and 11). Buckingham: Open University Press.

Parfitt AM. 2001. Skeletal heterogeneity and the purposes of bone remodelling: Implications for the understanding of osteoporosis. In: Marcus R, Feldman D, Kelsey J, editors. Osteoporosis, 2nd ed. San Diego, CA: Academic, pp. 433-447.

Parfitt AM. 2002. Misconceptions (2): Turnover is always higher in cancellous than in cortical bone. Bone 30(6):807-809.

Parker PL. 1964. The biogeochemistry of the stable isotopes of carbon in a marine bay. Geochimica et Cosmochimica Acta 28:1155-1164. Pate FD. 1994. Bone chemistry and paleodiet. Journal of Archaeological Method and Theory l(2):161-209.

Pendergast DM. 1992. Noblesse oblige: The elites of Altun Ha and Lamanai, Belize. In: Chase DZ, Chase AF, editors. Mesoamerican Elites: An Archaeological Assessment. Norman: University of Oklahoma Press, pp. 61-79.

Peterson BJ, Fry B. 1987. Stable isotopes in ecosystem studies. Annual Review of Ecology and Systematics 18:293-320.

Pfeiffer S, King P. 1983. Cortical bone formation and diet among Protohistoric Iroquoians. American Journal of Physical Anthropology 60:23-28.

Pfeiffer S, Varney TL. 2000. Quantifying histological and chemical preservation in archaeological bone. In: Ambrose SH, Katzenberg MA, editors. Biogeochemical Approaches to Paleodietary Analysis. New York: Kluwer Academic/Plenum Publishers, pp. 141-158.

Phenice T. 1969. A newly developed visual method of sexing in the os pubis. American Journal of Physical Anthropology 30:297-301.

Pinhasi R, Mays S. 2008. Advances in Human Palaeopathology. Chichester, West Sussex UK: John Wiley & Sons Ltd.

Pohl M. 1976. Ethnozoology of the Maya: An Analysis of Fauna from Five Sites in the Peten, Guatemala. Ph.D. Dissertation, Cambridge: Harvard University.

Pohl M. 1981. Ritual continuity and transformation in Mesoamerica: Reconstruction the ancient Maya Cuch Ritual. American Antiquity 46:513-529.

Pohl M. 1983. Maya ritual faunas: Vertebrate remains from five sites in Peten, Guatemala. In: Willey GR, editor. Excavations at Seibal, Department of Peten, Guatemala. Memoirs of the Peabody Museum of Archaeology and Ethnology, 15. Cambridge: Harvard University, pp. 147-174.

Pohl M. 1985. The privileges of Maya elites: Prehistoric vertebrate fauna from Seibal. In: Pohl M, editor. Prehistoric Lowland Maya Environment and Subsistence Economy. Papers of the Peabody museum of Archaeology and Ethnology, 77. Cambridge: Harvard University, pp. 133-145.

Pohl MD. 1990a. The ethnozoology of the Maya: Faunal remains from five sites in Peten, Guatemala. In: Willey GR, editor. Excavations at Seibal. Memoirs of the Peabody Museum of Archaeology and Ethnology, 17. Cambridge: Harvard University, pp. 142- 174. 194

Pohl MD. 1990b. The Rio Hondo Project in Northern Belize. In: Pohl MD, editor. Ancient Maya Wetland Agriculture: Excavations at Albion Island. Boulder: Westview Press, pp. 1-19.

Pohl MD. 1994. The economics and politics of Maya meat eating. In: Brumfield EM, editor. The Economic Anthropology of the State. Lanham, Maryland: University Press of America, pp. 119-148.

Pohl MD, Pope KO, Jones JG, Jacob JS, Pipemo DR, deFrance SD, Lentz DL, Gifford JA, Danforth ME, Josserand JK. 1996. Early agriculture in the Maya lowlands. Latin American Antiquity 7(4):355-372.

Pollard AM, Heron C. 2008. The chemistry of human bone: Diet, nutrition, status and mobility. In: Pollard AM, Heron C, editors. Archaeological Chemistry, 2nd Ed. Cambridge: Royal Society of Chemistry, pp. 346-328.

Pollock HED, Ray CE. 1957. Notes on Vertebrate Animal Remains from Mayapdn. Report No. 41. Washington D. C: Department of Archaeology, Carnegie Institution of Washington.

Powell ML. 1985. The analysis of dental wear and caries for dietary reconstmction. In: Gilbert RI Jr, Mielke JH, editors. The Analysis of Prehistoric Diets. Orlando, FL: Academic Press, pp. 307-338

Powis TG. 2004a. The role of pottery and food consumption among Late Preclassic Maya commoners at Lamanai, Belize. In: Lohse JC, Valdez F Jr, editors. Ancient Maya Commoners. Austin: University of Texas Press, pp. 49-72.

Powis TG. 2004b. Ancient lowland Maya utilization of freshwater pearly mussels (Nephronaias spp.). In: Emery KF, editor. Maya Zooarchaeology: New Directions in Method and Theory. Monograph 51. Los Angeles: UCLA Cotsen Institute of Archaeology,, pp. 125-140.

Powis TG, Stanchly N, White CD, Healy PF, Awe JJ, Longstaffe F. 1999. A reconstmction of Middle Preclassic Maya subsistence economy at Cahal Pech, Belize. Antiquity 73:364-376.

Price TD, Schoeninger MJ, Armelagos GJ. 1985. Bone chemistry and past behaviour: An overview. Journal of Human Evolution 14:419-447.

Price TD, Burton JH, Sharer RJ, Buikstra JE, Wright LE, Traxler LP, Miller KA. 2010. Kings and commoners at Copan: Isotopic evidence for origins and movement in the Classic Maya period. Journal of Anthropological Archaeology 29:15-32.

Prowse TL. 2001. Isotopic and Dental Evidence for Diet from the Necropolis oflsola Sacra (1st - 3rd Centuries AD), Italy. Ph.D. Dissertation, Hamilton, ON: McMaster University. 195

Prowse TL, Schwarcz HP, Garnesy P, Knyf M, Macchiarelli R, Bondioli L. 2007. Isotopic evidence for age-related immigration to imperial Rome. American Journal of Physical Anthropology 132:510-519.

Quade J, Cerling, TE, Barry JC, Morgan ME, Pilbeam DR, Chivas AR, Lee-Thorp JA, van der Merwe NJ. 1992. A 16-Ma record of paleodiet using carbon and oxygen isotopes in fossil teeth from Pakistan. Isotope Geoscience 94:183-192.

Redfield R, Villa Rojas A. 1962. Chan Kom, A Maya village. Chicago: University of Chicago Press.

Reed DM. 1994. Ancient Maya diet at Copan, Honduras, as determined through the analysis of stable carbon and nitrogen isotopes. In: Sobolik KD, editor. Paleonutrition: The Diet and Health of Prehistoric Americans. Center for Archaeological Investigations, Occasional Paper No. 22. Carbondale: South Illinois University, pp. 210-221.

Reed DM. 1998. Ancient Maya Diet at Copan, Honduras. Ph.D. Dissertation, State College, PA: The Pennsylvania State University.

Reed DM. 1999. Cuisine from Hun-Nal-Ye. In: White CD, editor. Reconstructing Ancient Maya Diet. Salt Lake City: University of Utah Press, pp. 183-196.

Rink WJ, Schwarcz HP. 1995. Tests for diagenesis in tooth enamel: ESR dating signals and carbonate contents. Journal of Archaeological Science 22:251-255.

Roberts CA, Manchester K. 2007. The Archaeology of Disease, 3r ed. Ithaca, NY: Cornell University Press.

Rose JC, Condon KW, Goodman AH. 1985. Diet and dentition: Developmental disturbances. In: Gilbert RI Jr, Mielke JH, editors. The Analysis of Prehistoric Diets. . Orlando: Academic Press, pp. 281-305.

Saul FP. 1972. The Human Skeletal Remains from Altar de Sacrificos: An Osteobiographic Analysis. Peabody Museum of Archaeology and Ethnology Paper 63(2). Cambridge: Harvard University.

Saul FP. 1975. Human remains from Lubaantun. In: Hammond N, editor. Lubaantun: A Classic Maya Realm. Monographs of the Peabody Museum of Archaeology and Ethnology, 2. Cambridge: Harvard University, pp. 389-410.

Saul FP. 1977. The paleopathology of anemia in Mexico and Guatemala. In: Cockbum E, editor. Porotic Hyperostosis: An Enquiry, Monograph No. 2. Detroit, MI: Paleopathology Association, pp. 10-15.

Saul FP, Saul JM. 1991. The Preclassic population of Cuello. In: Hammond N, editor. Cuello: An Early Maya Community in Belize. Cambridge: Cambridge University Press, pp. 134-158. 196

Saul JM, Saul FP. 1997. The Preclassic skeletons from Cuello. In: Whittington SL, Reed DM, editors. Bones of the Maya: Studies of Ancient Skeletons. Washington DC: Smithsonian Institution, pp. 28-50.

Saunders SR. 2000. Subadult skeletons and growth-related studies. In: Katzenberg MA, Saunders SR, editors. Biological Anthropology of the Human Skeleton. New York: Wiley Liss, pp. 136-161.

Saunders SR, Hoppa RD. 1993. Growth deficit in survivors and non-survivors: Biological mortality bias in subadult skeletal samples. Yearbook of Physical Anthropology 36:127- 151.

Scherer AK, Wright LE, Yoder CJ. 2007. Bioarchaeological evidence for social and temporal differences in diet at Piedras Negras, Guatemala. Latin American Antiquity 18(1):85-104.

Schoeninger MJ. 1989. Reconstructing prehistoric human diet. In: Price TD, editor. The Chemistry of Prehistoric Human Bone. Cambridge: Cambridge University Press, pp. 38- 67.

Schoeninger MJ, DeNiro MJ. 1983. Stable nitrogen isotope ratios of bone collagen reflect marine and terrestrial components of prehistoric human diet. Science 220:1381-1383.

Schoeninger MJ, DeNiro MJ. 1984. Nitrogen and carbon isotopic composition of bone collagen from marine and terrestrial animals. Geochimica et Cosmochimica Acta 48:625- 639.

Schoeninger MJ, Moore K. 1992. Bone stable isotope studies in archaeology. Journal of World Prehistory 6(2):247-296.

Schoeninger MJ, DeNiro MJ, Tauber H. 1983. Stable nitrogen isotope ratios of bone collagen reflect marine and terrestrial components of prehistoric human diet. Science 220(4604):1381-1383.

Schoeninger MJ, Moore KM, Murray ML, Kingston JD. 1989. Detection of bone preservation in archaeological and fossil samples. Applied Geochemistry 4:281-292.

Schoeninger MJ, Hallin K, Reeser H, Valley JW, Foumelle J. 2003. Isotopic alteration of mammalian tooth enamel. International Journal of Osteoarchaeology 13:11-19.

Schour L, Massler M. 1941. The development of the human dentition. Journal of the American Dental Association 28:1153-1160.

Schwarcz HP. 1991. Some theoretical aspects of isotope paleodiet studies. Journal of Archaeological Science 18:261 -275. 197

Schwarcz HP, Melbye J, Katzenberg MA,.Knyf M. 1985. Stable isotopes in human skeletons of southern Ontario: Reconstructing paleodiet. Journal of Archaeological Science 12:187-206.

Schwarcz HP, Dupras TL, Fairgrieve SI. 1999. 15N enrichment in the Sahara: In search of a global relationship. Journal of Archaeological Science 26:629-636.

Scott EC. 1979. Principal axis analysis of dental attrition data. American Journal of Physical Anthropology 51(2):203-211.

Scott RF IV. 1980..Further-comments on faunal analysis and ancient subsistence activities at Colha. In: Hester TR, Eaton JD, Shafer HJ, editors. The Colha Project: Second Season 1980. Interim Report. San Antonio: Center for Archaeological Research, University of Texas, pp. 281-288.

Sealy JC, van der Merwe NJ. 1986. Isotopic assessment of Holocene human diets in the southwestern Cape, South Africa. Current Anthropology 27:135-150.

Sealy JC, van der Merwe NJ, Lee-Thorp JA, Lanham J. 1987. Nitrogen isotope ecology in southern Africa: Implications for environmental and dietary tracing. Geochimica et Cosmochimica Acta 51 2707-2717.

Sealy JC, Morris AG, Armstrong R, Markell A, Schrire C. 1993. An Historic skeleton from the Slave Lodge at Vergelegen. Goodwin Series 7:84-91.

Sealy J, Armstrong R, Schrire C. 1995. Beyond lifetime averages: tracing life histories ' through isotopic analysis of different calcified tissues from archaeological human skeletons. Antiquity 69:290-300.

Seidemann RM, McKillop H. 2007. Dental indicators of diet and health for the Postclassic coastal Maya on Wild Cane Cay, Belize. Ancient Mesoamerica 18:303-313.

Shannon IL, Carroll EC, Madsen KO. 1970. Dietary influences on the formation of dental calculus in rats. Journal of Periodontal Research 5:191-195

Sharer RJ, Trazler LP. 2006. The Ancient Maya. Stanford: Stanford University Press.

Sheets P. 2000. Provisioning the Ceren household: The vertical economy, village economy, and household economy in the southeastern Maya periphery. Ancient Mesoamerica 11:217-230.

Sheiham A. 1983. Sugars in dental decay. The Lancet 1:282-284.

Shemesh A. 1990. Crystallinity and diagenesis of sedimentary apatites. Geochimica et Cosmochimica Acta 54:2433-2438.

Shennan S. 1988. Quantifying Archaeology. Edinburgh: Edinburgh University Press. Siemens AH, Puleston DE. 1972. Ridged fields and associated features in southern Campeche: New perspectives on the Lowland Maya. American Antiquity 37:228-239.

Sillen A. 1989. Diagenesis of the inorganic phase of cortical bone. In: Price TD, editor. The Chemistry of Prehistoric Human Bone. Cambridge: Cambridge University Press, pp. 211-229.

Sillen A, Sealy JC. 1995. Diagenesis of strontium in fossil bone: A reconsideration of Nelson et al. (1986). Journal of Archaeological Science 22:312-320.

Singh S, Potturi BR. 1978. Greater sciatic notch in sex determination. Journal of Anatomy 125(3):619-624.

Smith AL. 1972a. Excavations at Altar de Sacrificios: Architecture, Settlement, Burials and Caches. Peabody Museum of Archaeology and Ethnology Papers Vol. 62. Cambridge: Harvard University.

Smith BH. 1984. Patterns of molar wear in hunter-gatherers and agriculturalists. American Journal of Physical Anthropology 63:39-56.

Smith BN. 1972b. Natural abundance of the stable isotopes of carbon in biological systems. BioScience 22(4):226-231.

Smith BN, Epstein S. 1971. Two categories of 13C/12C for higher plants. Plant Physiology 47(3):380-384.

Smith RE. 1955. Ceramic sequence at Uaxactun, Guatemala (2 vols). Middle American Research Institute, Publication 20.

Song R. 2004. Reconstructing Infant Diet and Weaning Behavior of Ancient Maya from Lamanai, Belize Using Laser Ablation-Inductively Coupled Plasma-Mass Spectrometry (LA-ICP-MS). Ph.D. Dissertation. Amherst: University of Masschusetts Amherst.

Sponheimer M, Lee-Thorp J A. 1999. Alteration of enamel carbonate environments during fossilization. Journal of Archaeological Science 26:143-150.

Stafford TW Jr, Brendel K, Duhamel RC. 1988. Radiocarbon, 13C and 15N analysis of fossil bones: Removal of humates with XAD-2 resin. Geochimica et Cosmochimica Acta 52:2257-2267.

Steggerda M. 1941. The Maya Indians of Yucatan. Publication no. 531. Washington, DC: Carnegie Institution of Washington.

Stewart TD. 1974. Human Skeletal Remains from Dzibilchaltun, Yucatan, Mexico, with a Review of Cranial Deformity Types in the Maya Region. Middle American Research Institute, Vol. 31. New Orleans: Tulane University, pp. 199-225. Stewart TD. 1979. Essentials of Forensic Anthropology. Springfield, IL: Charles C Thomas.

Steyn M, Iscan MY. 1997. Sex determination from the femur and tibia in South African whites. Forensic Science International 90(1): 111-119.

Stress B. 2010. This world and beyond: Food practices and the social order in Mayan religion. In: Staller JE, Carrasco M, editors. Pre-Columbian Foodways: Interdisciplinary Approaches to Food, Culture, and Markets in Ancient Mesoamerica. New York: Springer Science, pp. 553-576.

Stuart-Macadam P. 1985. Porotic hyperostosis: Representative of a childhood condition. American Journal of Physical Anthropology 66:391-398.

Stuart-Macadam P. 1988. Nutrition and anaemia in past human populations. In: Kennedy BV, LeMoine GM, editors. Diet and Subsistence: Current Archaeological Perspectives. Calgary: The University of Calgary Archaeological Association, pp. 284-287.

Stuart-Macadam PL. 1989. Nutritional deficiency disease: A survey of scurvy, rickets and iron-deficiency anemia. In: Iscan MY, Kennedy KAR, editors. Reconstruction of Life from the Skeleton. New York: Wiley-Liss, pp. 201-222.

Stuart-Macadam P. 1992. Anemia in past human populations. In: Stuart-Macadam P, Kent S, editors. Diet, Demography and Disease: Changing Perspectives on Anemia. New York: Aldine de Gruyter, pp. 151-170.

Stuart-Macadam P, Kent S, editors. 1992. Diet, Demography and Disease: Changing Perspectives on Anemia. New York: Aldine de Gruyter.

Suchey J, Katz D. 1986. Skeletal age standards derived from an extensive multiracial sample of modem Americans. Abstract. American Journal of Physical Anthropology 69:269.

Sullivan CH, Kruger HW. 1981. Carbon isotope analysis of separate chemical phases in modem and fossil bone. Nature 292:333-335.

Sullivan CH, Kruger HW. 1983. Carbon isotope ratios of bone apatite and animal diet reconstmction. Nature 301:177.

Tayles N, Domett K, Nelsen K. 2000. Agriculture and dental caries? The case of rice in prehistoric Southeast Asia. World Archaeology 32(l):68-83.

Tieszen LL. 1991. Natural variations in the carbon isotope values of plants: implications for archaeology, ecology, and paleoecology. Journal of Archaeological Science 18:227- 248. 200

Tieszen LL, Boutton TW. 1989. Stable carbon isotopes in terrestrial ecological research. In: Rundel PW, Ehleringer JR, Nagy KA, editors. Stable Isotopes in Ecological Research. Ecological Studies 68. Berlin: Springer-Verlag, pp. 167-195.

Tieszen LL, Fagre T. 1993. Effect of diet quality and composition on the isotopic composition of respiratory CO2, bone collagen, bioapatite, and soft tissue. In: Lambert JB, Grape G, editors. Prehistoric Human Bone: Archaeology at the Molecular Level. Berlin: Springer-Verlag, pp. 121-155.

Todd TW. 1921a. Age changes in the pubic bone, I: The male white pubis. American Journal of Physical Anthropology 3:285-334.

Todd TW. 1921b. Age changes in the pubic bone, III: The pubis of the white female, IV: The pubis of the female white-negro hybrid. American Journal of Physical Anthropology 4:1-70.

Tozzer AM. 1941. Landa's Relacion de las Cosas de Yucatan. Papers of the Peabody Museum of American Archaeology and Ethnology, vol. 18. Cambridge: Harvard University.

Trautz OR. 1967. Crystalline organization of dental mineral. In: Miles AED, editor. Structural and Chemical Organization of Teeth, Vol. 2. New York: Academic Press, pp. 165-200.

Triffitt JT. 1980. The organic matrix of bone tissue. In: Urist MR, editor. Fundamental and Clinical Bone Physiology. Philadelphia: J. B. Lippincott, pp. 45-82.

Troughton JH, Wells PV, Mooney HA. 1974. Photosynthetic mechanisms and paleoecology from carbon isotope ratios in ancient specimens of C4 and CAM plants. Science 185(4151):610-612.

Turner BL II. 1978. The development and demise of the swidden thesis of Maya agriculture. In: Harrison PD, Turner BL II, editors. Pre-Hispanic Maya Agriculture. Albuquerque: University of New Mexico, pp. 13-22.

Turner BL II, Harrison PD. 1981. Prehistoric raised-field agriculture in the Maya Lowlands. Science 213(4506):399-405.

Turner BL II, Miksicek CH. 1984. Economic plant species associated with prehistoric agriculture in the Maya Lowlands. Economic Botany 38(2): 179-193.

Tuross N, Fogel ML, Hare PE. 1988. Variability in the preservation of the isotopic composition of collagen from fossil bone. Geochimica et Cosmochimica Acta 52:929- 935. 201

Tuxill J, Reyes LA, Moreno LL, Uicab VC, Jarvis DL 2010. All maize is not created equal: Maize variety choices and Mayan foodways in rural Yucatan, Mexico. In: Staller JE, Carrasco M, editors. Pre-Columbian Foodways: Interdisciplinary Approaches to Food, Culture, and Markets in Ancient Mesoamerica. New York: Springer Science, pp. 467-486.

Tykot RH. 2002. Contribution of stable isotope analysis to understanding dietary variation among the Maya. In: Jakes KA, editors. Archaeological Chemistry: Materials, Methods and Meaning. Washington, DC: American Chemical Society, pp. 214-230.

Tykot RH. 2006. Isotope analyses and the histories of maize. In: Staller J, Tykot R, Benz B, editors. Histories of Maize: Multidisciplinary Approaches to the Prehistory, Linguistics, Biogeography, Domestication and Evolution of Maize. Burlington, MA: Academic Press, pp. 131-142.

Tykot RH, van der Merwe NJ, Hammond H. 1996. Stable isotope analysis of bone collagen, bone apatite, and tooth enamel in the reconstmction of human diet: A case study from Cuello, Belize. In: Oma MV, editor. Archaeological Chemistry: Organic, Inorganic and Biochemical Analysis. ACS Symposium Series No. 625. Washington, DC: American Chemical Society, pp. 355-365.

Ubelaker DH. 1978. Human Skeletal Remains. Washington, DC: Taraxacum Press.

Urey HC. 1947. The thermodynamic properties of isotopic substances. Journal of the Chemical Society 1947:562.

Valadez R. 1999. Los animals domesticos. Arqueologia Mexicana 6(35):32-39.

Vander Zaden MJ, Cabana G, Rasmussen JB. 1997. Comparing trophic position of freshwater fish calculated using stable nitrogen isotope ratios (S15N) and literature dietary data. Canadian Journal of Fisheries and Aquatic Sciences 54:1142-1158.

Vanderklift MA, Ponsard S. 2003. Sources of variation in consumer-diet <515N enrichment: A meta-analysis. Oecologia 136:169-182. van der Merwe NJ. 1982. Carbon isotopes, photosynthesis and archaeology. American Scientist 70:596-606. van der Merwe NJ, Medina E. 1989. Photosynthesis and 13C/12C ratios in Amazonian rain forests. Geochimica et Cosmochimica Acta 53:1091-1094. van der Merwe NJ, Medina E. 1991. The canopy effect, carbon isotope ratios and foodwebs in Amazonia. Journal of Archaeological Science 18:249-259. van der Merwe NJ, Vogel JC. 1978. C content of human collagen as a measure of prehistoric diet in woodland North America. Nature 276:815-816. van der Merwe NJ, Rosevelt AC, Vogel JC. 1981. Isotopic evidence for prehistoric subsistence change at Parmana, Venezuela. Nature 292:536-538.

van der Merwe NJ, Tykot RH, Hammond N, Oakberg K. 2000. Diet and animal husbandry of the Preclassic Maya at Cuello, Belize: Isotopic and zooarchaeological evidence. In: Ambrose SH, Katzenberg MA, editors. Biogeochemical Approaches to Paleodietary Analysis. New York: Kluwer Academic/Plenum Publishers, pp. 23-38.

van der Veen M. 2003. When is food a luxury? World Archaeology 34(3):405-427.

van Klinken GJ. 1999. Bone Collagen Quality Indicators for Palaeodietary and Radiocarbon Measurements. Journal of Archaeological Science 26(6):687-695.

VanPool TL, Leonard RD. 2011. Quantitative Analysis in Archaeology. Maiden, MA: Wiley-Blackwell.

Villa Rojas A. 1945. The Maya of East Central Quintana Roo. Publication 559.Washington, DC: Carnegie Institution of Washington.

Vogel JC. 1978. Isotopic assessment of the dietary habits of ungulates. South African Journal of Science 74:298-301.

Vogel JC. 1993. Variability of carbon isotope fractionation during photosynthesis. In: Ehleringer JR, Hall AE, Farquhar GD, editors. Stable Isotopes and Plant Carbon-Water Relations. San Diego: Academic Press, pp. 29-46.

Vogel JC, van der Merwe NJ. 1977. Isotopic evidence for early maize cultivation in New York State. American Antiquity 42:238-242.

Wada E, Kadonaga T., Matsuo S. 1975.15N abundance in nitrogen of naturally occurring substance and global assessment of denitrification from isotopic viewpoint. Geochemical Journal 9:139-148.

Walker PW, DeNiro MJ. 1986. Stable nitrogen and carbon isotope ratios in bone collagen as indicators of prehistoric dietary dependence on marine and terrestrial resources in southern California. American Journal of Physical Anthropology 71:51-61.

Walker PL, Dean G, Shapiro P. 1991. Estimating age from tooth wear in archaeological populations. In. Kelley M, Larsen CS, editors. Advances in Dental Anthropology. New York: Alan R. Liss, pp. 169-178.

Walker PL, Bathurst RR, Richman R, Gjerdrum T, Andrushko VA. 2009. The causes of porotic hyperostosis and cribra orbitalia: A reappraisal of the iron-deficiency-anemia hypothesis. American Journal of Physical Anthropology 139:109-125. 203

Wang Y, Cerling TE. 1994. A model of fossil tooth and bone diagenesis: Implications for paleodiet reconstruction from stable isotopes. Palaeogeography, Palaeoclimatology, Palaeoecology 107:281-289.

Wapler U, Crubezy E, Schultz M. 2004. Is cribra orbitalia synonymous with anemia? Analysis and interpretation of cranial pathology in Sudan. American Journal of Physical Anthropology 123:333-339.

Waters-Rist AL, Katzenberg MA. 2010. The effect of growth on stable nitrogen isotope ratios in subadult bone collagen. International Journal of Osteoarchaeology 20:172-191.

Webb EA, Schwarcz HP, Healy PF. 2004. Detection of ancient maize in lowland Maya soils using stable carbon isotopes: evidence from Caracol, Belize. Journal of Archaeological Science 31:1039-1052.

Webb EA, Schwarcz HP, Jensen CT, Terry RE, Moriarty MD, Emery KF. 2007. Stable carbon isotope signature of ancient maize agriculture in the soils of Motul De San Jose, Guatemala. Geoarchaeology: An International Journal 22(3):291.-312.

Webster JW, Brook GA, Railsback LB, Cheng H, Edwards RL, Alexander C, Reeder PP. 2007. Stalagmite evidence from Belize indicating significant droughts at the time of Preclassic abandonment, the Maya hiatus, and the Classic Maya collapse. Palaeogeography, Palaeoclimatology, Palaeoecology 250:1-17.

Weiner S, Bar-Yosef O. 1990. States of preservation of bones from prehistoric sites in the Near East: Survey. Journal of Archaeological Science 17:187-196.

White CD. 1986. Paleodiet and Nutrition of the Ancient Maya at Lamanai, Belize: A Study of Trace Elements, Stable Isotopes, Nutritional and Dental Pathologies. Master's Thesis, Peterborough, ON: Trent University.

White CD. 1988. Diet and health in the ancient Maya at Lamanai, Belize. In: Kennedy BV, LeMoine GM, editors. Diet and Subsistence: Current Anthropological Perspectives. Calgary: University of Calgary Archaeological Association, pp. 288-296.

White CD. 1997. Ancient diet at Lamanai and Pacbitun: Implications for the ecological model of collapse. In: Whittington SL, Reed DM, editors. Bones of the Maya: Studies of Ancient Skeletons. Washington, DC: Smithsonian Institution Press, pp. 171-180.

White CD, editor. 1999. Reconstructing Ancient Maya Diet. Salt Lake City: The University of Utah Press.

White CD. 2004. Gendered food behaviour among the Maya: Time, place, status and ritual. Journal of Social Archaeology 5(3):356-382.

White CD, Armelagos GJ. 1997. Osteopenia and stable isotope ratios in bone collagen of Nubian female mummies. American Journal of Physical Anthropology 103(2): 185-99 White CD, Schwarcz HP. 1989. Ancient Maya diet: As inferred from isotopic and elemental analysis of human bone. Journal of Archaeological Science 16:451-474.

White CD, Healy PF, Schwarcz HP. 1993. Intensive agriculture, social status, and Maya diet at Pacbitun, Belize. Journal of Anthropological Research 49(4):347-375.

White CD, Wright LE, Pendergast DM. 1994. Biological dismption in the early Colonial period at Lamanai. In: Larsen CS, Milner GR, editors. In the Wake of Contact: Biological Responses to Conquest. New York: Wiley-Liss, pp. 135-145.

White CD, Pohl MD, Schwarcz HP, Longstaffe FJ. 2001a. Isotopic evidence for Maya patterns of deer and dog use at Preclassic Colha. Journal of Archaeological Science 28:89-107.

White CD, Pendergast DM, Longstaffe FJ, Law KR. 2001b. Social complexity and food systems at Altun Ha, Belize: The isotopic evidence. Latin American Antiquity 12(4):371- 393.

White CD, Pohl MD, Schwarcz HP, Longstaffe FJ. 2004. Feast, field and forest: Deer and dog diets at Lagartero and Copan. In: Emery K, editor. Maya Zooarchaeology: New Directions in Method and Theory. Monograph 51. Los Angeles, CA: UCLA Institute of Archaeology, pp. 141-158.

White CD, Maxwell J, Dolphin A, Williams J, Longstaffe F. 2006a. Pathoecology and paleodiet in Postclassic/Historic Maya from northern coastal Belize. Memorias do Instituto Oswaldo Cruz 10 (Suppl. II): 1-9.

White CD, Longstaffe FJ, Schwarcz HP. 2006b. Social directions in the isotopic anthropology of maize in the Maya region. In: Staller J, Tykot R, Benz B, editors. Histories of Maize: Multidisciplinary Approaches to the Prehistory, Linguistics, Biogeography, Domestication, and Evolution of Maize. Burlington, MA: Academic Press.

White CD, Price TD, Longstaffe FJ. 2007. Residential histories of the human sacrifices at the Moon Pyramid, : Evidence from oxygen and strontium isotopes. Ancient Mesoamerica 18:159-172.

Whittington SL. 1999. Caries and antemortem tooth loss at Copan: Implications for a commoner diet. In: White CD, editor. Reconstructing Ancient Maya Diet. Salt Lake City: University of Utah Press, pp. 151-168.

Whittington SL. 2003. Analysis of human skeletal material excavated by Guillemin. In: Nance CR, Whittington SL, Borg BE, editors. Archaeology and Ethnohistory of Iximche. Gainesville: University Press of Florida, pp. 241-304.

Whittington SL, Reed DM. 1997. Commoner diet at Copan: Insights from stable isotopes and porotic hyperostosis. In: Whittington SL, Reed DM, editors. Bones of the Maya: Studies of Ancient Skeletons. Washington: Smithsonian Institution Press, pp. 157-180. 205

Whittington SL, Reed DM. 1998. Evidencia de dieta y salud en los esqueletos de Iximche. Mesoamerica 35:73-82.

Willey GR. 1974. The Classic Maya hiatus: A "rehearsal" for the collapse? In: Hammond N, editor. Mesoamerican Archaeology: New Approaches. London: Duckworth, pp. 417- 430.

Willey GR, Bullard WR Jr, Glass JB, Gifford JC. 1965. Prehistoric Maya Settlements in the Belize Valley. Papers of the Peabody Museum of Archaeology and Ethnology, vol. 54. Cambridge: Harvard University.

Williams JS. 2000. The People Who Ate the Sea: A Stable Isotope Analysis of Diet at Marco Gonzalez and San Pedro, Belize. Master's Thesis, London, ON: The University of Western Ontario.

Williams JS, White CD. 2006. Dental modification in the Postclassic population from Lamanai, Belize. Ancient Mesoamerica 17:139-151.

Williams JS, White CD, Longstaffe FJ. 2005. Trophic level and macronutrient shift effects associated with the weaning process in the Maya Postclassic. American Journal of Physical Anthropology 128:781-790.

Williams JS, White CD, Longstaffe FJ. 2009. Maya marine subsistence: Isotopic evidence from Marco Gonzalez and San Pedro, Belize. Latin American Antiquity 20(l):37-56.

Williams PL, Warwick R, Dyson M, Bannister LH, editors. 1989. Gray's Anatomy, 37th ed. Edinburgh: Churchill Livingstone.

Wing ES. 1975. Appendix 6: Animal remains from Lubaantun. In: Hammond N, Lubaantun: A Classic Maya Realm. Monographs of the Peabody Museum of Archaeology and Ethnology, no 2. Cambridge: Harvard University, pp. 379-383.

Wing ES. 1981. A comparison of Olmec and Maya foodways. In: Benson E, editor. The Olmec and Their Neighbors. Washington, DC: Dumbarton Oaks, pp. 21-28,

Wing ES, Brown AB. 1979. Paleonutrition: Method and Theory in Prehistoric Foodways. New York: Academic Press.

Wing ES, Scudder SJ. 1991. The exploitation of animals. In: Hammond N, editor. Cuello: An Early Maya Community in Belize. Cambridge: Cambridge University Press, pp. 84-97.

Wing ES, Steadman D. 1980. Appendix: Vertebrate faunal remains from Dzibilchaltun. In: Andrews EW IV, Andrews EW V, editors. Excavations at Dzibilchaltun, Yucatan, Mexico. Middle American Research Institute Publication No. 48. New Orleans: Tulane University. 206

Wright ACS, Romney DH, Arbuckle RH, Vial VE. 1959. Land Use in British Honduras: Report of the British Honduras Land Use Survey Team. Colonial Research Publications 24. London: The Colonial Office.

Wright LE. 1994. The Sacrifice of the Earth? Diet, Health, and Inequality in the Pasion Maya Lowlands. Ph.D. Dissertation, Chicago, IL: The University of Chicago.

Wright LE. 1997a. Ecology or society? Paleodiet and the collapse of the Pasion Maya Lowlands. In: Whittington SL, Reed DM, editors. Bones of the Maya: Studies of Ancient Skeletons. Washington, DC: Smithsonian Institution Press, pp. 181-195.

Wright LE. 1997b. Biological perspectives on the collapse of the Pasion Maya. Ancient Mesoamerica 8:267-273.

Wright LE. 2005. Identifying immigrants to Tikal, Guatemala: Defining local variability in strontium isotope ratios of human tooth enamel. Journal of Archaeological Science 32:555-566.

Wright LE. 2006. Diet, Health, and Status among the Pasion Maya: A Reappraisal of the Collapse. Vanderbilt Institute of Mesoamerican Archaeology Vol. 2. Nashville: Vanderbilt University Press.

Wright LE, Chew F. 1998. Porotic hyperostosis and paleoepidemiology: A forensic perspective on anemia among the ancient Maya. American Anthropologist 100(4):924- 939.

Wright L, Schwarcz HP. 1996. Infrared and isotopic evidence for diagenesis of bone apatite at Dos Pilas, Guatemala: Palaeodietary Implications. Journal of Archaeological Science 23:933-944.

Wright LE, Schwarcz HP. 1998. Stable carbon and oxygen isotopes in human tooth enamel: Identifying breastfeeding and weaning in prehistory. American Journal of Physical Anthropology 106:1-18.

Wright LE, Schwarcz HP. 1999. Correspondence between stable carbon, oxygen and nitrogen isotopes in human tooth enamel and dentine: Infant diets at Kaminaljuyu. Journal of Archaeological Science 26:1159-1170.

Wright LE, White CDr 1996. Human biology in the Classic Maya collapse: Evidence from paleopathology and paleodiet. Journal of World Prehistory 10(2):147-198.

Wright, L. E., J. A. Valdes, J. H. Burton, T. D. Price, and H. P. Schwarcz. 2010. The children of Kaminaljuyu: Isotopic insight into diet and long distance interaction in Mesoamerica. Journal of Anthropological Archaeology 29:155-178.

Ximenez F. 1967. Historia Natural del Reino de Guatemala. Socidedad de Geografia e Historia, Publication Especial 14. Guatemala City: Editorial Jose de Pineda Ibarra. 207

Zazzo A, Lecuyer C, Mariotti A. 2004. Experimentally-controlled carbon and oxygen isotope exchange between bioapatites and water under inorganic and microbially- mediated conditions. Geochimica et Cosmochimica Acta 68(1):1-12.

Zimmerman DW, Zumbo BD. 1993. The relative power of parametric and nonparametric statistical methods. In: Keren G, Lewis C, editors. A Handbook for Data Analysis in the Behavioral Sciences: Methodological Issues. Hillsdale, NJ: Lawrence Erlbaum Associates, pp. 481-517. 208

APPENDIX A:

Non-Human Skeletal Remains and Artefacts 209

Burial #1, Str. A-1

A nondiagnostic reddish-grey ceramic fragment and two nondiagnostic plaster fragments were identified amongst the fragmented remains from Burial #1. In addition, several long bones were identified that did not conform to any human bones; one was identified as an ungulate long bone, although the others could not be more specifically identified. Finally, a minimum of 6 freshwater mussel shell fragments, possibly from

Nephronaias ortomanni, were identified. The shells had been perforated at one end, suggesting that they had been strung together to form a piece of jewellery.

Burial #4, Str. C-2

Within the nondiagnostic remains from Burial #4 two human long bone fragments and one human rib fragment that may have exhibited cut marks were identified. It is possible, however, that these marks represent postmortem damage. Furthermore, two small bone fragments, possibly belonging to a bird (species unknown) were identified.

One bone fragment had been worked and exhibited an incised line across one end and a perforated hole in the centre. It is unknown whether this bone is human or non-human..

Burial #5, Str. C-l

A minimum of three rib fragments belonging to CI-5-A exhibited what looked to be cut marks, although it is possible that these marks represent postmortem damage. The only artefact included in the burial was a ceramic body sherd (Awe 1985:115), fragments of which were identified among the skeletal remains. The three recovered pieces were refitted, and had a dark (grey-black) outer layer, and where the sherd was fragmented a reddish-orange temper was visible.

Unknown Burial 210

Among the skeletal remains whose provenience was unknown a non-human bone fragment and several freshwater mussel shells were identified. The non-human bone is certainly an articular surface, and the bone resembles a human glenoid cavity but is far too circular and deep to be from a human skeleton. A more specific identification could not be made. Finally, a complete freshwater mussel (possibly Nephronaias ortomanni) shell was identified. It was smaller than the shells recovered from Burial #1 and was not perforated. 211

APPENDED B:

Plant and Animal Remains Recovered from Ancient Maya Sites Plant Species Recovered from Ancient Maya Sites

Common Name Scientific Name Site(s) Time Period(s)* Source(s) \IHIA: (com) Zea mays Cerros LPC Crane (1996); Crane and Carr (1994) : Pulltrouser Swamp Beginning in A Pohl et al. (1996) Caracol EC,LC Healy etal. (1983) ,, "!"'". '' CahalPech LPC Powis etal. (1999) Cuello EPC " ".",", Miksicek etal. (1981) /. , V|ii.^li Cucurbita sp. Cerros LPC Crane (1996); Crane and Carr (1994) • Cuello" PRC , Hammond and Miksicek (1981) i, Cahal Pech LPC Powis etal. (1999)

lloan ; Phaseolus vulgaris Cerros ,wJr\W' Crane (1996); Crane and Carr (1994) ' ( ullnii Gossypium cf. Hirsutum Cerros LPC Crane(1996) (. liili pepper Capsicum $p..: Cerros LPC "" .' " /•"" Crane(1996)" " "'7' ( ;iv..ii< Theobroma cacao Pulltrouser Swamp EC, LC Miksicek (1983); Turner and Miksicek (1984) Copan EC,LC Ilammdnd and Miksicek (1981 j; Turner and < • Miksicek (1,984) ,; Vmce Byrsonima crassifolia Cerros LPC Crane (1996); Crane and Carr (1994) n^mmmm^ Pulltrouser Swamp EC,LC Miksicek (1983); Turner and Miksicek (1984) Copan EC,LC Turner and Miksicek (1984)

• ..!;• '.', Cuello PRC Hammond and Miksicek (1981); Turner and fcfs^islS-?i!l'' "'A • Miksicek (1984) Coyol palm Acromia Mexicana Cerros LPC Crane (1996); Crane and Carr (1994) : ,v:i Mamey Calocarpumtnammosum Cerros LPC Crane (1996); Crane and Carr (1994) Pulltrouser Swamp EC, LC Miksicek (1983); Turner and Mjksicek (1984). Cuello "" PRC "'.';•' Hammond and MJksicef;(198T); turner and ..? ". .•:..i.-..:.' • •?f,' ••= ->;i: Miksicek 91984). .;:,.'•' :Cf : Guava Psidium cf. Guajava Cerros LPC Crane(1996)

V Pulltrouser Swamp tlV^9 \J\Jt \ UMiksiSeEp83^'Turner anlMksiwk:(|984)' ] Cuello PRC Hammond and Miksicek (1981); Turner and Miksicek (1984) ! : SniLoie Cor dig. cf Dodecandra ; ...Cerrds' .•-' "•"•"""'LP^."-' .J; V- ,Crah?(1996); Crane and Carr (19943? ,| Pulltrouser Swamp EC,LC Miksicek (1983); Turner and Miksicek (1984) Plant Resources Continued Common Name Scientific Name Site(s) Time Period(s)'' Source(s) J fl^^'Sjbulj, ... r ^"-' -yMastichodendron sp. Cirros T ,;. fi (^©!|1996TrCr4ne^Mdi^arljCl'994),,??.;%' • Persimmon Diospyros sp. Cerros 1_11LPCT_^_ Crane(1996) ,pf^j^y^aao$.;:. ;".' .. Persea-type Cerros - %;. Pulltrouser Swamp ^EC,LC' Miksicek (J1983); Turner and Miksicek (1984) Copan '",'. • ^Htomohl'&d Milli|^^Bi); "fitner^d. J

Cuello _^__ * PRC _ __ TurneIan(* Miksicek (1984) Papaya Caarica -type Cerros ''-"T" Pulltrouser Swamp _ ECjLC"" Miksicek (1983);jrurner~and Miksicek (1984) ; :. licara Crescentia - type Cerros '. .T'vLPC'l'7T P '~'~J~W: Crate^9'6j ' ~";. " Caimito Chrysophyllum sp. Cerros ...wLPCj"_ Crane(1996) llouplum Spondiassp. Cuello '.: MikliceE(1991); HamiiiiMul and Miksicek .v-.':&;.• ' (198 n Pulltrouser Swamp ^ EC,'LC_ Miksicek (1983); Turner and Miksicek (J984)_ MKpiee : r igmriPimenla sp:.-,..''. Cuello ;,' £?& Miklp^991) " " _•-• i Pulltrouser EC, LC Miksicek (1983); Turner and Miksicek (1984) Ca(If§sl:tfee -^JQrescentid sp. Pulltrouser '" F £C;LC '" MiIsice||l983);:Tfi«iSnd Miksicek(i984)" SapodiUa Manilkara zapota Pulltrouser EC, LC Miks]ceHl983JlJurner and Miksicek (1984) Cuello !V r HammondlandPMiUuvki l'»8l i. Inn j^ " PRC""' { : ^ ; ;,,MikMeekillJ8-I| Copal Protium copal Copan Turner and Miksicek (1984) Cerros EC, LC -™„ . ©SBand'l^^liy^r ""••£ Jaucate palm Bactris sp. Cuello Hammond and Miksicek (1981); Turner and PRC Miksicek (1984) ^T^Chrysophyllum sp, Cueii.i I kmimoml MiidMlllicekf P'X I). 'I urner ;uid lllipat-'apTle^ PRC Manioc Manihot esculenla Cuell" I l'( Miksicek (3""11 T^n^^'^j^if^ Pulltrouser Swimip llcuinning in A. Cattail Typha sp. Cerros LPC Crane(1996)

to u> Mammal Remains Recovered from Ancient Maya Sites

Common Name Scientific Name Site(s) Time Period(s)* Source(s) White-tailed deer Odocoileus virginianus Caledonia LC Awe j: 1985) Cerros LPC Carr (1986) , • Altar de Sacrificios MP*C,LPC;MC,LC Pohf(1983a) Seibal MPC, LPC, MC, LC Pohl(1983a; 1983b)

, , l-'V^j Zl«/w I\^ 5< > ' Tikal ( 9 JoUlm^u Macanche TLC, PC Pohl(]98-;ii

pc +.-. -' , ' , (1 * FI0S5 .' Pohl(198 vu Cuello EPC, LPC Pohl(198«bi -,'"•"' HopS! EC Pohl (1.983 hi Uaxactun LC Pohl(198«hi Copan ""LC' ~" "" Pohl(1983b) Mayapan PC Pohl(198«hi

8 ' . ' ' Cozuraet "PC' Pohl(198-h! Aguateca LC Emery (2'H ni < , ' ' ' Arrofi de Piedra EC, MC Emery (2<» 10) Tamarind ito EC, LC Emery (2'»l (M --*----—-~-~~-~"~r "' CabalSJech' LPC Powis et\'f (ll""») TLC, PC Emery (2" MM ^ " — Ma|po Gonzalez PC Seymour (IW1) Quim Chi Hi Ian PRC Emery (2d Id) i . ' Peccary Tayassmdae sp. Cozumel Island PC ' " HamblinHl>8li: [VhlH'Whi Cerros 1 P( Carr (198''i: ( nine :m>| i .uiil,»,»li Seibal MPC. 1 P( .1 (' Pohl (198 hi: !1»XMD • Altar de Sacrificios l( Pohl (198 VI) * "Tikal IK . PC Pohl (198 vi: IWh> Mammal Remains Continued Common Name Scientific Name Site(s) Time Period(s)'* Source(s) Peee;n\ /./Ic/SSf'/i/i/C sp Macanche \ TLCPC./< Pdhl (I983#i|;^ Flores PC Pohi"(1983a)~"U* Copan .' ;,cT. - '•rr Uaxactun LC Pohl(1983b)

sMayapan j?:Pohl,(^83b) '. Dzibilchaltun "LC * Pohl(1983b) I men |20 10) "LC I men. C 2d IH) :Cerros •'•;"' •"*-;' C an (l«>Xfii Altar de Sacrificios " "MPC, LPC, EC, LC"™ P.'hl(ll»Svi) Domclic l)n» ('win \lilll'l.iri Dos Pilas l( Emery (201 m Aguateca K" I may"(2])1P),. Arroyo dePiedm K ,K I meiy (2010) iCahal Pech I P( • ' P.'wis^/vA(199Q) Punta de Chimino TLC Emery (2IIIIM ! Qu|||j '%Z'$W r-*%fit "7Emery(20l(M Seibal MPC, LPC, EC,YC Pohl (1983aTl983b) Mammal Remains Continued Common Name Scientific Name Site(s) Time Period(s)* Source(s) : Domestic Dog ^•y^Ganis familiar i> Tikal ' - "•=«:;.. 'H.TLC,PCv::; '^-...' #&jtp983a) »•• .'•; Brocket deer Mazama sp. Macanche TLC, PC " 'pohT(1983a) :i?:f;K PohJ(f9i3a) •J Aguateca LC 1 mery(2010) •r.: ""1C. 11 I iWyiSllOR •• ,.vi;' , - £ • •? : V Cahal Pech LP( Powis etal. (1999) SJ' TamarittdiS /( \.r.rX^: Seibal MI' .1 l*( . K Pohl(1983a) r.'.v, ".„ . Tikal ,;t; . "T; i£* • n Pc|I|S83a).;gr^ *.;-;• • . j^rj-j ' 'possum Didelphis spp. Holmul EC Pohl (1983b)*"" Mayapart".':-""'-' '•'; •;.:' ;.PC -.;;.:'.' gf '.^ f»-iii (i ox' i»f : •• Altar de Sacrificios MPC, LPC, LC P.»hhl,»X,M) r : , - -,:?,' . ; "ify- -M piM' ;•;•""~i ""LC'T "'71' '''"^'r'pi.|il(l«)X?:i) Piiin.i Felis concolor Tikal TLC, PC Pohhl'iX'Mi Zr-M-;- Flores Seibal LC Pohl (1983a) Dos Pilas .,... j.LC . E5e^(5S|(r)v -^ Jaguar Felis onca Aguateca ' LC Emery"(2010) Altun Ha^ : . ,'6:i"' •- . •.:' "'I'-y-- '-v.-' I;EC :'pM(pi6>rj^" San Jose LC Pohl (1983b) Manatee, ^'X-Tfl^h'1 hit* iihiinttus Altun Ha; V\ ': JSj T|p|j^9Slb|rf^r~' San Jose LC "Pohl (1983b) "" Freshwater Resources Recovered from Ancient Maya Sites

Common Name Scientific Name Site Time Period(s)* Source(s) Musk Turtle Staurotypus spp. ||jjltar de Sacrificios;;:',;. '•' MPC/LPCyLC,-: • Pohl (1983a). "H, •#] Seibal __ MPC, LPC, LC _ Pohl(1983a) *\Klacanehe V;;r:•'..';• ' T- " X JLA-fi JT V/ -,'f Pohl (198.i;n Pond Turtle Pseudemys scripta Altar de Sacrificios MPC, LPC, LC" "PohI(198*iu , Seibal "-;— ' "'"-.:-^ • • MPC, LP^OJ ..,$$,- "Pohl("1983,1. IWbi Tikal _„_.„ TLC, PC Pohl (198 ';II : MacancheJ;: 'JLJL/Vrf'* Jpv**'' •'„. Pohl (1983a) Flores PC Pohl(19XM> ; ; Mud Turtle Kinosternon spp. Seibal ' ";'^r'"- """Mpc.ipc^icir"" :....:^Pihl'(tf8 :i) Altun Ha EC Pohl (198 ?.i) Dos Pilas LC " " Emeryfioioi - Aguateca LC Emery (2<> MM Arroyo de Piedras MC ' ' ""•'•" ' Emery (20 UK Punta de Chimino TLC, PC Emery (2n 1 "I Black-bellied turtle Rhinoclemmys areolata Altar de Sacrificios MPC,LF€ T" ' Poh?(198vii Tikal TLC, PC Pohi (1983a) Crpcodile Crocodylus spp. Altar deSiKiifiui's \1P( . 1 P( . I ( l»ofir(i983i)^-.: -"w~>A Seibal \1P( IP' 1 ( Pohl(1983a; 1983b) Macanchc IK . PC ifJIPJlBa) ' ..'.^'I'i'.-^^l Tikal PR<)« Pohl(1983b) -,'','_,' - Tulum II ( l^|l983h)^^X:v ;iJC : Dzibilchaliun II ( Pohl(1983b) Aguateca K Punta de< liiiiinii> IK P( 'PflBPl

Marine Resources Recovered from Ancient Maya Sites

Common Name Scientific Name Site(s) Time Period(s)* Source(s) Shark Squattfopaes Mayapan PC Pohl|1983B)' :••, Piedras Negras PoW(1983b) , Palenque '• A-. **,, '""' I'JttCj; "Pohl (1983b)'- "?T" Altun % ^ ~ ' EC " _ _ jPohl (1983b) ^~?j: «^ ' •" "j-wfija^ii,^ CerrosT .A-A^ . XPC"~" ".-"7"" ' Craneted Carr (I?p4) Siiii'ji,i\ |).i«\.iud.ic I axactiin __ LPC,EC,LC Pohl(1983b) V!s likal " " '" AT ;FoM(t983b> • •«& Aj ' opan LPC Pohl(1983b) \ltard© Sacrificios :p^l(ip3b)''V'"7v' \ltun Ha EC Pohl(1983b) loniGtta ...,;;., • • • JbCj l/iu. ..$•• ;foMtl983b)!iA ' I 'zibilchaltun LC Pohl(1983b) _ C ozumel Islands "PolFtW^^A ^..'l Mayapan __ PC Pohl(1983b) lii-iieiMi llhnlii vhl/>c\ CahalPech ' v '/|. *' iP0w&^'d,':(199p|",' "A|| Cerros __ _^ LPC Crane and Carr (1994) ^"" slacks ip:.Qarangidai^. CerrbSA • AA 1A • ^LPCl;'""''1; Cran^an|i?|iir..(1994). |f>. ^ Jacks Carangidae Marco Gonzalez Postclassic Seymour (1991) crab >- Catimectes sapidm CerrosJr. . ^-.^r '•' " LKT"-~" ; Crane, and CH(i994K,, • • Al to 00 Marine Resources Continued Common Name Scientific Name Site(s) Time Period(s)* Source(s) Uarracuda Sphyracna sp. (Vims ....•it.fi''' »UN Lid; . • - Jv-• r- AAI>: •"•'•-•".-v-Crane attd'Caif|1994)A - 1 Parrot Fish Sparisoma sp. Cahal Pech LPC Powis et al. (1999) Sea turtle Chelonioidea Cerros.^A . \-A'C: '• \W--t&F ..-#" -A^" . ,!S^elniCarR(:1994):AA5.-l Sea Snails Melungea sp. Cerros LPC Crane and Carr (1994) : : V Oliva sp. Do, Pilas TK " " ,- ,;'7SeryWfdFl ;:v A ''lJ \jll.ltCi..l K Emery (2010) \IIHVU Je Piedra 1 < /^A»ffiMM-l§A^:ii ( ili.il PeJi IP' Powis et al. (1999) 'Juiin I 111 Milan I' - ;::fEMery'|Sfo)v%-.. AA-A Conch Slrombus sp. ' em". IP' Crane and Carr (1996) D.vJ'il.is _ 1 ( , 1 C. K _A...lJ::ip%lP^2lS....'lll"^ Cahal Pech LPC Powis etal. (1999) Aguateca fL,;,A AV . ^.^^ ^;;;;-ji^;:ns^plo)''ASA^"T'^ Spondylus sp. Dos Pilas LC Emery (2010)

Avian Resources Recovered from Ancient Maya Sites Common Name Scientific Name Site(s) Time Period(s) Source(s) l (Kcilhiial 1 Hikes !/••/. • ',177 i»i; Hiila Seibal 1 ( Pohl(l')S3a: l >X3b) hkal I K . P( PMIII(I'»X^II \1aa>. anehc II C. P'' l\>hl(ll>X3a) 1 lines P( Pi-hlil'iX^ii t (Viimel MC Pi«hlil,»8',iij 1 ilium IK Pi'liliP'X;;!! \iiu.itcea 1 ( I.mei> (201(11 ( ah.il Peeh 1 P< PnwiNi/.// ( I""1)) Oscillated Turkey \ini\i< de PieJras l.( 1 mer> (20101 Puma Je ( liimiiii' II ' 1 mei\ I2II|UI Pa i rot el. llHJZ niw /'/' Seibal I.C PnhidyX'iM Tikal TLC, PC Pohl (1983a) *See key provided below for time period designations. 220

Key for Time Period Abbreviations: A Archaic EPC Early Preclassic MPC Middle Preclassic LPC Late Preclassic PRC Preclassic PROC Protoclassic EC Early Classic MC Middle Classic LC Late Classic TLC Terminal Late Classic PC Early Postclassic C Undefined Classic period 221

APPENDIX C:

Floral and Faunal Stable Carbon and Nitrogen Isotope Data Used to Create Food Web Plant Stable Carbon and Nitrogen Isotope Values

Resource Species Modern Corrected/Archaeological w Source Category 8 CCQi i sL 8 CCQi' 2a Msii/i- Zi'ii /'/iir.s 40.1 -8.6 10.0* Keegan and DeNiro (1988) / "Ml 1 , .^AL1-2 -9.7 6.1 W right (2006:95) /. mm s f -9.5 3.6 Wright \2M:95) / mm \ -8.2 Norr (1991:134) /. H/iM * -9.6 -8.1 1.8 Norr (1^91:134) / H/.M \ -10.1 -8.6 Norr (1991:134) f / ///./IS -9.8 -8.3 ."'" J 1.4 Norr(1991:f34) Seagrasses Thalassia testudiunum (turtle grass) -6.2 -4.7 0.9 Keegan and DeNiro (1988) Syrmgodiumjfiliforme (manatee grass) -12.8 4L3 A ' 2.0 .Keegan and DeNiro (1988) Macroalgae Laurencia sp. (brown alga) -15.7 -14.2 6.2 Keegan and DeNiro (1988) Laurencia sp. (bnfipatga)' -15.7 -14.2 o: KLMUMI and DeNiro (1988) Batophera sp (green alga) -5.8 -4.3 i" kieyan and DeNiro (1988) Hahmeda sp. (calSRous green alga) 42 40,5 -» -* kee;j^in,'d D'elSfird f§l8) Unidentified encrusting red alga -14.3 -12.8 _ ^ kee-jan and DeNiro (1988) .. C3. plants Z>/o5cfflB0KH&^nl 'r ' -25.1 -23.6 ~(> Ksvjgnd DeNiro (1988) -26.3 Dioscorea spJL(yam) -24.8 ^ ki-Lsam and DeNiro (1988) A::'L:' ';A Diosca^^^^^mj -27 -25.5 ^ •• k .vjjpfnd DeNiro (1988) Dioscorea sp. (yam) -27.3 -25.8 :s kee-an and DeNiro (1988) DioscbreSTsp. (yantjr' -28.6 -21A i« Wofr (1991:134) ", Dioscorea alata (macal/yam) -25.1 -23.6 - Wright (2006:95) Manihot esculenta (manioc) -27.6 -26.3 2 \ kccs-ati and DeNiro (1988) M. esculenta (manioc) -26.1 24.6 Norr (1991:134) , Unidentified tuber (Eddde) -24.4 -22,9 ;. •:" 3i. '"I ieegaSandbeNiro(1988) Ipomaea batatas (sw. Potato) -25.7 -24.2 3.8 Keegan and DeNiro (1988) -26l4 -24.9 - Wright (2006:95) ... Opuntia sp. (nopal cactus) -io.o -8.5 - Wright (2006:94) Capsicumsp, (chili pepper) -30.1 -28.6 _ , ,- flight (2006:94) Capsicum sp. (duhjepper) -27.8 -26.3 - Wright (2006:94) \Brn\j mum iihk l(ramon) -27.7' -26.* \\ii^u(:n()6>n) Plants Stable Carbon and Nitrogen Isotope Values Continued :T37 Corrected/Archaeological Plant Species Modern 8"CC0, Source Category m 613Cco, (%o) J%o)_ C3 Plants B. alkastrum W right (2006:94) Continued B. alicastrum W light (2006:95) Aslrocaryum mexicanum (chapay nut) -i i ,-i -2*).l» »:?4) A. mexicanum -;n.l W right (2006:94) 1 , Orbignya colune (corozo nut) . -2<».~ -2S.2 ST'" O. colune -:s' -2i. S W light (2006:94) Pouteria mamosa (zapote) -2". I -25.(1 W rigSfScppf^ P. mamosa -:~x ll.h W right (2006:94) P. mamosa -28.: -2(v7 W right (2006:95);^ P. mamosa -30.0 -28.5 Wright (2006:95) Bixa orellana (achiote) • -29.7 -28.2 "Wright (20|ipfl' Annona muricata (soursop) -27.7 -26.2 Wright (2006:94) Artnona sp. -29.0 -27.5 Wright (2(fp3P" Caricasp. (wild papaya) -26.4 -24.9 Wright (2006:94) I'SA-V'*' . jft^^^^icacao (cacao) -34.1 -32.6 "Wright(20^5) -.i Dialium guianense (wild tamarind) -26.1 -24.6 Wright (2006:95^ _ 'Ps^^-j^^ava (guava) -27.3 -25.8 Wnghf(2C Lincania platypus (sunzapote) -29.3 -27.8 .Wright (2006j95) si A A' ' ClA^P^^ip. (pepitoria squash) -27.4 -25.9 88" \\ri hi (2C Cucurbita sp. -27.1 -25.6 W i ml" (2006:95) '. ByriS^^crassifolia (nance) -27.1 -25.6 W i idii (20HP)' B. crassifolia -28.3 -26.8 \<>\i 11991:134) i ,Zjr ^'".A'.Sw^^P1 leaf) -26.7 -25.2 10 \\n^hi'(20W5)V (waterhly) -24.6 -23.1 \\imht(2006:95y ;•;' ; A Ba^^^ipaes (peach palm) -27.1 -25.6 : i Non (19^ JSP).. Legumes Phaseolus vulgaris (beans) -27.1 -25.6 Wright (2006:94f -27.8 -26.3 3.9 _ P. vuhn^s -28.3 -26.8 0.5 N(OT(1991:134); "*~ -28.6 -27.1 Phaseolus lunatus (lima bean) -28 -26.5 -2.2 Norr(1991:134f" """' "Removed from sample because value was outside expected parameters. to to Terrestrial Resource Stable Carbon and Nitrogen Isotope Values

Corrected / Modern 8"C:13/ i 815N Resource Category Species co Archaeological Source (%o) 13 (%o) 8 CCQi (%o) 1 t'lioslriiil .Mmiiiiiiil \ I, iii inn i. inwricana (Red brocket -21.6 4.7 White and Schwarcz (1989) deer) \d. Americana -22 4.3 White and Schwarcz (1989) ^^^piericana 20.4 4.3 WhimS (1993)" ?A \i. Americana •24.3 3.7 Williams etal. (2009) • 19.9 8.1 WilliaM^^(2009)-':; Odocoileus virginianus (White- •23.7 15.8 White er al. (1993) i ailed deer) i). Vk'ginimtm •21.6 7 While i7.i/. (199TJT l). I'irginianus •19.5 5 White etal. (1993) O. VWJIfhianus .'v/7' •17.5 10.8 White etal. (1993)-r l). Virginianus •13.5 9.4 White etal. (1993) a rfflfan* 7' :' •19.8 6.2 Willii§g|fer a/. (2009T i." O. I'irginianus •19.3 4.6 Williams etal. (2009)_ (). Virginianus •18.5 2.3 N(^99i:;l30)Ar:,; i (). I'irginianus •21.2 3.6 Norr (1991:130) O. vWginianus'-• {••.,,'/ •20.6 3.5 Norr (1991:1301^'";.• O. I'irginianus •19.7 Norr (1991:130) (I Virginianus •20.8 3.2 Nc>rr099i:Yl3O)A"" \. (). I'irginianus -20 2.5 Norr(1991:130) O. Virginianus •29.2 4 Norr {1991: l5b).'"';'ii (). Virginianus •22.2 3.2 Norr (1991:130) (). Vir^imanus. •21.7 2.8 Norr (jo«» 1.1 W) Iiipirus baindii(Tapir) •23.3 4.9 While and Scliwaie/i I'Wi , I'I'WIWH layacu (Collared peccary) •13.5 7.8 White tv.//. ilv >3) / /.M./nV .->->-> 3.2 Norr(l'»"l Mm /\i\\[:n>ciapunctata (Agouti) •21.2 4 Norr (1991 130) to to 4^ Terrestrial Resource Stable Carbon and Nitrogen Isotope Values Continued 513/ Corrected / 15 Modern 8"C „i 8 N Resource Category Species c Archaeological Source (%») 13 (%o) 6 Ccol(%o) Terrestrial Mammal -lOf 8 9 White and Schwarcz (1989)^ Continued -8.: * Whiter al. (1993) -19.1 5 X WMiaSlR>i§lb09) "" •19.5 -18 h tCeegan and DeNiro (1988) l.iguaiiq -19.' 5.h "ANorr(199ill!)). • f. iguana -21." ; D ' Norr (1991 M30)___ Qm§0 -21.' i •» "Norr(f99i:;130)""" "* l. iguana -|9.o ^ S Norr (1991:130) •X •» Ifrrc^trial Crab I nkimmicuh -23.8 jy^ \ jgeegan and Deljlg'(1988J'"' Freshwater Resource Stable Carbon and Nitrogen Isotope Values

Corrected / Resource Modern 15 Species 13 Archaeological 8 N (%o) Source Category 8 CC„, (%o) 13 8 CCQi (%o) Freshwater Fish Coiorada fish -28.6 -27.1 9.2 Wrigto(2J)06:W)l) Al~ Colorada fish -27.4 -25.9 10 _ Wright |2006:101) Guapote fish -30.8 , -29.3 9.3 Guapote fish -29.9 -28.4 11.5 __ Wru^tC2006:10J\_ ' Pelenia splendida (Bay snook) -29.5 "' -28 ' 12.3- P. splendida (Bay snook) -32.7 -31.2 13.3 Wrighf(2006:101) /*. splendida (Bay snook) -29.6 -28.1 ,, Il;9 ,, " Wg|j^Hpl01> ' ; P. splendida (Bay snook) -32.1 -30.6 11.6 WrightttOOfcioi) _ Ictglurus sp. (Catfish) -22.3 "" -20.8" "" ',-11.59 Pachychilus glaphyrus (Jute snail) Wright (2006:101) -33.6 -32.1 4.9 meat P. glaphyrus (Jute snail; meat) ' -30 '.'-28.5 A- "JR P. glaphyrus (Jute snail; shell) -13.8 -12.3 - __ ^_Wright^0^10l£ ''•¥: glaphyrus (Jute'snail; shell) •4"3.'JT '"'"•42.3.. A >:: "jA'-- Amblema sp. (Freshwater mussel; -11.4 -9.9 - Wright (2006:101) shell) Amblema sp. (Freshwater mussel; -12.4 -10.9 - Wrtgl1(20p6dyl|r:"'' '\ shell) Fn-tliwater Crab Unknown Crab -21.2 -19.7 3.21 Wright (2006:ToY) ^ Kiver Reptile; Unknown Turtle ,' -21.4 4.1 White IJBJfwap: (I9$9j] Unknown Turtle _->-> -> 5.3 White and Schwarcz (1989) Dermatemys mawii (river turtle); -20.2 5.6 WilHW^"<2009)""" '" Trachemys scripta (pond turtle) -19.5 4.3 Williams er al. (2009) C 'r9XS^ sp> (crocodile) ' -18.2 8.9 Wiljigmsirfaf. &Q09) Marine Resource Stable Carbon and Nitrogen Isotope Values

Corrected/ Modern 813C i Species co Archaeological 81SN (%o) Source (%o) 13 ' 8 CC„ (%o) Haemulon flavolineatum (yellow p uiui -7 -5.5 5.7 Kees-Mii mid DeNtee (1988) \ H. flavolineatum -7 -5.5 5.3 keegan and DeNiro (1988) H. flavolineat^£ -8.2 -6.7 5.4 Keegan md DeNiro (1988X4 ~"% //. «a/ra(sailor^s choice) ' -7.5 -6 5.1 keegan and DeNiro (1988) " - -4.8 4.6 VffllP at (2009)'^' 1 Lutjanus ahogany (hog snapper) -8.5 -7 4.3 keegan and DeNiro (1988) . I„Bii§nutton snapper)' -4.7 -3.2 6.4 KccP^HB^^^j|p8)^ } L. jocu (dog snapper) -8 -6.5 6.6 keegj^nd^eNiro^SS)^^ ___ jgjBaMKitfir ^idnc snapper} -6.4 -4.9 5.7 Kccl&^^^e^^^^pSji-. ' j Trachinotus sp. (Pompano) - -6.2 16.6 Williams etal. (2009) ; Acanthurus spf|i|irgeon fish) ;.;,. • -I0.1 ^g-TT ''"""• Williams etal (2009) ^ Ocyurus chrysurus (yellowtail snapper) -7.7 -6.2 6.1 Keegan and DeNiro (1988) ^^^^^^MMtf^^KiiW^^^Wf' ' A . -9.4 -7.9 5.7 Keegan and DeNiro (1988) Priacanthus cruentatus (glasseye snapper) -I2.l -10.6 5.6 Keegan and DeNiro (1988) ffaemulo/i(gnmt);X: "'•''. :3A "'"'"^' • -4.9 6.6 Keegan and DeNiro (1988) Sparisoma (parrotfish) - -7.8 3.9 Keegan and DeNiro (1988) k &^^^^g»Otfish>;/.-;.;; ;--.:'' -, - -5.6 8.8 Williams "et al (2009) Calamus (porgy) -8.2 9.2 Keegan and DeNiro (1988) Caranl&ubet QaciMsh) .7*4... • -1.3 8.1 K ccgan and DeNiro (1988) A| C. ruber - -1.3 8.3 Keegan and DeNiro (1988) : Caranxsp. \'A^ /'....k^'rA-' •- - -4.4 10 wmmSetat, (2009)T" "^| Caranx sp. - -5 12.4 Williams et al. (2009) Batistes vetttla (queen triggerfish) -7.1 -5.6 5.9 KccgmiandDeNko (1988) Balistes sp. (trigger fish) - -7 7.1 Williams el al (2009) .' Arhts sp, (river catfeh). . - : .Ac-. - -6.4 11.2 millmSetd (200$) Heimramphiis sp. (halfbeak) -14.I -12.6 8.4 Keegan and DeNiro (1988) BSS^^afii ^*;^- •."'•"'•:<:'' '"''"] -14.3 -12.8 8.5 Keegan and,DeNko||J^,88)' A|J Albula vulpes (bonefish) - -3.8 -1.4 Keegan and DeNiro (1988) to to -J Marine Resource Stable Carbon and Nitrogen Isotope Values Continued Corrected/ Modern 8 13/ Species ^col Archaeological 815N (%o) Source (%o) 813Cc„, (%o) A. vulpes -M 5 koL,Mii .Hid neNiro()988) 1 1 Ginglymostoma cirratum (nurse shark) -" 1 in kLCLMii .mil DLNfiro (1988) -8 1 S kircLMii .nidi )eNiro (1988)' ; Carcharhinus (requiem shark) -»); in i k«.eLi.iii .nid De\firo(1988) Aeipbatis narmari (spotted eagle ray54, -ion 10.5 KLC.MII.HIII DeNiro (1988)4" \ A. narinari -1 ! * 42 KLL-^III .nid DeNiro (1988) Sphyracna barracuda (barracuda) 4 ' -'9 S^ Rix-uan .mil l)eNird^988) " ") S. barracuda -3.8 8.2 Keegan and DeNiro (1988) S. barracuda -5.9 11.1 Williams etal (2009) ] Euthynnus (tuna) -14.6 l-.S keegan and DeNiro (1988) Epinephetus guttatus (Nassau grouper)- 41.6 40.1 ""^ • keegan and DeNiro (1988) j E. striatus (red hind) -8.4 -6.9 '0 keegan and DeNiro (1988) Mycteroperca (grouper) -8,1 8." Williams¥^(2009) j Mycteroperca (grouper) -6.7 m: Williams "et al."(2009) l^^^^^^^^rapper) ,:"" ""-5,3 ~.9 Wl!|iaml^^^^H > Chondrichthyes (sharks) -9.2 11.8 Wiiiiamsera/. (20091 """4-3,8 4"" '"X "':41i ' ! Keegan an(|^®l)fo (19XX i Unidentified encrusting black sponge _-5.4 -3.9 3.9 keegan and DeNiro (J9XX1 ^•A~^^^Srbonaria (sponge) , ""..'. -15'"" '43X" "447^-"~.-'Ci .^ "Keegan aacfjllpiro (19XX) Chondrilla nucula (sponge) -13.3 -11.8 5.2 Keegan and DeNiro (19XXi •12S WfcV ' Keegan' anc^^^^|8S i Condylactis gigantea (anemone) -11.5 -10 4.8 Keegan and DeNiro (198J0 ^^5^tifiMA«r^|^^^^^p"y4' ' -123 . . . ,^£,^_£.. 71^3.84 " Tripneustes esculentus (urchin) -4.5 -3 2.5 Keegan and DeNiro (1988) " T. EsculenfmWT^ '"' ""'^',",..."."'"'" '""; ."?• :$.i """•""" . " Z3""" Keegan attd^^^^^^Ei) Diadema antillarium (black urchin) -9.3 -7.8 2.4 KeeganandDeNiro (1988) Unidentified Holothuriidae -13 -11.5 6.3 Keegan and DeNiro (1988)'"""'

00 Marine Resource Carbon and Nitrogen Isotope Values Continued Corrected/ 13 15 Species Modern 8 CC0, (%o) Archaeological 8 N (%o) Source 13 8 CC„, (%») ^^limkmmrium (gastropod)' -10.9 , -9.4 3.74 •'(• Keegan and DeNiro (1988) ] Atrina rigida (pelecypod) -13.4 -11.9 2.3 Keegan and DeNiro (1988) PliMB^ (pelecypod) . -9.8 ''>-... •""4C:> ' -8.3 '."2.6 44 T ';"'4«1feegan:andp^|3^88) •','• Nerita versicolor (littoral gastropod) -5.7 -4.2 2.1 Keegan and DeNiro (1988) ''{Mlvefsicolor • -"X -ii '< 2.9 Keegan and'b»6(f988) ' N. versicolor -l-X -i i ; % ^ keegan and DeNiro (1988) t;M^Si--Sqiiaftti''iii^ (lilloial aiiipliiiieiiiail) -X " -d" ^ ; Keegan ^SlRiro (198 8) J C. squamosus -X" -~ i ; ^ keegan and DeNiro (1988) iS^^amasus -8 9 -~ ! !.l Keegan jjliSgira (1988) Strombus giga^ iine'j;i'j;isiinp'idi -i:n -II i : i keegan and DeNiro (1988) 1 C^^^aprbknUiris (pelce\ pod) -2^ < -20 8 keegan SBViro (l?!Pi: C. orbicularis -2' X -22 ' 11> Keeuan and DeNiro (1988) CMmicularis -V K -21 ? o 1 Keegan 5£f SiffiSR |988*

to to 230

Summary of Resource Category Means used to Construct Food Web

15 8 Ccoi 8 N Category 1S 8 Ccoi i („,) S.D. N 8 N (%») S.D. N %3:MatltS- i, •* ..., Maize -8.7

Sea grass and algae IferrestrM Resopr^.aM',

Land Crab I ie>hwjtef¥ih • A I ic-diujier Snail Meat Rccfl siuaitteFisl Inshore Fish d&horeflsh -r AA? Marine Invertebrates M-irmclSSS^drFT"^ Note: S.D. represents standard deviation and N represents number of sampled individuals 231

APPENDIX D:

Mean Stable Isotope Data for Ancient Maya Skeletons Mean Stable Isotope Data for Ancient Maya Skeletons

U 15 13 13 Site Time Period* 8 CC0, (%o) 8 N (%o) 8 Cbio(%o) 8 Cen (%») A Cbio-coi (%») Reference(s) Belize , - - % Altun Ha PRC 43.941.5(7) 10.610.7(8) -9.0±0.8 (7) - 4.811.0(7) White et al. (2001b) EC 40.1±1J (7) 10.6±0.4(7) -6,1±1J(8) ' 4.1=2.1 (7) Whiteero/.(2b01,b) " "\ LC -12.6±0.8(20) 10.3±0.3(19) -8.8±0.7(21) - 4.1*0.7(19) White etal. (2001b) " TC., , 4J.9±1.8(5) i0.4±0.4(5) ' -8.2±i.0(6) 3.910.6(5) White et at (2001b) " , PC -13.7(5) 10.7(5) -9.7 (5) 4.1(5) White etal. (2001b)

Baking Pot LC '41.0*1.0(9)' ~ 9.2±1.3(9) "-6.640.6 (4) - ' • 4.6H.I (4) *" Gerry (1993) "'" j Barton Ramie EC 41.4±0.9(7) 8.6±0.5 (7) -7.410.6 (6) - 4.210.9(6) Gerry (1993) Cahal Pech PRC 41.7±1..5(ll)" 9.4±Ll(il) «,' fj;r -» - 1*owis et at (1999) " 4 LC 41.211.5(31) 8.9±0.4(31) -6.9±0.5 (24) - 4.710.9 (23) Gerry (1993) , , Caledonia LPC 41.8(1) 8.3 (1) '" -7.7(1)"" ' -5.8 (1) 4.2(1) 1 his Study 4~~" EC,LC -9.1 ±0.4 (3) 8.4±0.2 (3) -4.910.4 (3) - 4.210.1 (3) This Study LC -10.2±2.1 (13) 9.0±0.8(13) -6.5±l.5(U) -3.611.3(4) 3.811.5(11) This Study ' 4 i Caracol EC -9.1 (8) 9.0 (8) - - - A. Chase and D. Chase (2001); A. Chase et al. (2001) l.C -10.0(72) 9.5 (72) - - - A. Chas£SDS^S£e (2001): A ciajiliooi) II C -10.8(5) 9.6(5) - - - A. Chase and D. Chase (2001); A Chase et al. (2001) Chau HII\ U -10.814.4(2) 10.7±0.5 (2) -7.4±0.2 (2) -6D ' 1- 16(2) MetcaI1^^^p)09a) "44^ K 42.710.9(2) 11.010.7(2) -8.0*1.3(2) -5(1. 1 " II Ii2) Metcalfe et al. C2009a) is III -11,8^-2.1(15) - 10.9±O?9 (15) ^8J±1.0(15) -6.2 ( In) .0.~(>) MetcaIf^Pg3gJ09"a) 444 (iiillu l'l« 42.910.9(28) 8.9*1.0(23) -9.8±1.0(16) -8.712.«( "I 1 1 Tykot et al. (1996); \ in der Merwe et al. (2000) K .i\oh MI'l -14.7±0.5(4) 9.3±0.3(4) -9a±0.5<5) - ft l) ' ( 1) Headerscma998s:2()03) 'j^ 1 l'( 44.5±0.8(4) 8.9±0.7 (4) -9.6±1.2(9) Ift II >(2> Henderson (1998, 2003) ( -15.5*1,0(2) 9>3±0.4 (3) -9.4±0.8 (9) ' ft 'i 0 1 i: i lletideimm(1998,3O03) 44 Lamanai PRC -12.7 (2) 10.2 (2) -6.9±0.2 (3) -8.8(1) 5.8T Coyston et al. (1999); White and Schwarcz (1989)

to to Maya Isotope Studies Continued 13 15 I3 I3 Site Time Period* 8 Ccol (%») 8 N (%o) 8 Cbi0 (%») 8 Cen (%o) A Cbio.col (%») Reference(s) 1 .im.-inai If ;-113±l.4(4)-' 10,9±1.3 (4) p =' -6.4*1,4 (4)' ••• " ,: . __ ~»'A / f! > .. = Cdys»f^Ml99^);Whtte,,A' ; ' "^ ^^^£^t "' ivm'd I .',; . 1 ',"'.•.:.- ...4'"'• > . ' ':.r'. . ;" ' A-- MSSSSSSiSS^l^ *.,:•> - ^ "'uan^PPiz (1989)"' -/A. _ 4|j 1 ( -14.110.9(3) 10.4(2) -6.110.8(4) - 8.6* "*"" Coyston et al (1999); White jind Schwarcz (1989) ! __.^ .._„„ II ( 4i5.6±i;f(7)' "•; ~9'MbA(if'-"-':"""-7.4iT.2'(4)" ' ""••••jfgfoyr t Coyston eHHSSilffp^'^ w& BcW^^^^M PC -9.310.8(24) 9.510.5 (24) -6.411/7 (18) -2.0(2) 2.9f Coyston et al. (1999); White and Schwarcz (1989); White et al. (1994) _.„ ,,_ .„_ : | rrr ' ^5*$6 (11) ' 49J±0.6(10) -5.610.7 I1') -I.Sdi v" l 1'WHII i' ji 1 I'N'M; \\ hik- ;iin| j ! •:•.'.'--••'.f'r ^ ' Vlmnu/ i llWM. \\ Imo . / oi (1"41) Marco PC,H "" -7.511.2(37) 10.5H.1 (37) -6.1H.2 (^"i - 1 1 II.Xi'A) WillianiMV.i/ (:i»il»i Gonzalez 4.-8,5i4W;?) ll/.in LC -9.210.3 (5) .8.010.9(5) - - - Wright (1994) kimilic PC -8.0(13) ^m^.Am:^-^i^M^m^^^^mi. . 'A*4;sV'"- • - .• ^SA4.II«)'A ^:\-T7^Pif^ •:iA..SA,',^^^&(2003) ' t.' „ to Maya Isotope Studies Continued 15 13 13 l3 Site Time Period* S^Ccol (%o) 8 N (%o) 8 Cbl0 (%o) 8 Cen (%») A Cbl0.C0, (%o)" Reference(s) K;iniiit4Mku . EC- -9.911.0(24), 9.2±i.8(15) 3^3.0*1* (96). „ Wright Mdfehwardz ( V. "'S i IVdras EC. -9.211.4(11) 8.310.6(11) Scherer et al. (2007) \cgras "'A'SfAj 'LC :' ''49116:8(34)""' 8.910 J|34) "ScfcriW'El^"!'"! vibal PRC -9.610.9 (7) 9.710.7 (7) """ Wright (1994) _ :. f^j -|0.2±0.5(3) 1O.0±&9(3) ^6,410^0)4^-' --W ' 3'."8Ti'.f(3p ^""Qi^limM. LC -9.611.3(24) 9.610.7 (37) -6.21(^9(24) 3.411.5(24)^ Gerry (1993); Wright (1"" 1' r:'tLC"' '".-9'4±I,i(i3) 8.6103(13) Wri| 1 :i\acttin LC _-10.7ll.O (6) 9.410.9 (5) \...:il

13 13 Mean 8 Ccoi and 8 Cbio Values for Ancient Maya Skeletons

I5 1J 1J Site 8"Ccoi(%») 5 N(%o) 8 Cbio (%o) 8 Cen (%o) \liun I hi -11 8 111 7 -8 } Klklll^ Pol -11 l>2 -h (i Barton Ramie -ii: XS -7 4 ( jhal Pech -11 5 " 2 -Ol) CaL-doni.i -ID x^; -h ( iiku.nl -10 l>4 - Chiiu 1 lik -I0.S lll.fi -7.5 l ( UL'llii -12" X > -»JX K'ii\oh -14.8 9 1 -'M 1 illllillKII -1 1 n W -<» 5 Marco (ion/ale/ _"• .i 10 5 -(* 1 \lolu> I ,|\ -S * 104 PIILIMUUI -102 9 1 -5 3 San Pedro -6.5 9.9 -4.3 XLIII.IICCII -l>.0 lM - \ll.u do Siicnlicioi -lM- So1 - DONPIKIN -v 1- W(,- - 1 lulllllll -w4 ') ; -4 * It/iin -1' 2 8 - l\ inn. lie -8 8 -

KilllllllilljllVLI ->).D 9 2 - I'lcdl.is V'L'l.ls -v: SA - Scikil -IJ ~ ») s- -fi."

I ilN.IClllll -in ~ "4 -•^ ~ C np.ui -in.: -.ii -^ *, ( liimv.hiiv.mil 44' ~ --4 ^ ilMlllil -i: -i — ^ -3 2 *Mean for several time periods calculated by present author. 236

APPENDIX E:

Porotic Hyperostosis and Dental Disease in the Caledonia Skeletal Sample Porotic Hyperostosis

Bone Side Location Observation tf! :, :: ||^^^^-';';i*eA!>*'%4?34^"'?L''':':' '': •a'rf^'';'4 '''4%!44* l- •• * "44444 M .ii ^""2.0.2,J2.03'"""*"' "'3 """"'"' 4" "(6.1.2, 6.2.5,6.33 7 r fipg^-A4— • 'T*:" ' ~; "TT.' • t nr^T"'. 2 0.2,2.0.3 3 4 6.1.2,6.2.5,6.3.3 "See Buikstra and Ubelaker (1994) for descriptions of pathology recording codes.

Linear Enamel Hypoplasia

Quadrant Tooth Number of Defects L /&&'', 'i I oft Maxilla Canine 2 1< i ght Maxilla, Can i ne •2' I eft Mandible Canine 1 l< i ;jht Mandible Canine 1

(I-5-A v l< 1 :-dit Maxilla I irst Incisor 2" Second Incisor 1

^ v'«' ,i*^'a'Kt/.„ - * Left Maxilla First Incisor •y msm^-f^^^ v^A^-'-Secpai Canine 3 238

Antemortem Tooth Wear, Dental Caries, Abscesses and Calculus

C2-3-A Qiudraut 1 (t()111 Presence DevelopmenDcM-lopn t Wear (siries Abscess ( alculus \1.I\III.II\ Ki-hi \H ^ - \12 •* \eM 0 u I B Ml ^ V.I4 (1 (1 IH. i 1 i

P: 1 V.I4 (1 0 1 11 ^ PI \LI 1 (1 II 1 Ii t ^ \A 1 0 II 2 I i i: 1 \v.l4 (1 II n i 1 \v.l 1 II 0 (i M.IMII.IIV I ell II 12 ^ ( V.I 1 ^ l) I) 1 It. 1 PI \v.N 1 0 (1 i r> P2 VI 1 1 li II i it

\1 \e!1 ;n 0 (i II \I2 \H M.iiiihhiil.il I ell \H \, 1 11 II ti (1 \12 \e 1 20 0 0 111

Ml \v. 1 2(. (i II 1 11

P2 \v. -1 1 0 II 2 1 i

PI \L 1 0 II ii ( 12

M iiidilnil.n ki'jln II 12 •i "

( •^ - PI 1 \L 1 1

P2 1 \v. 1 I

T M \s." 1 1 M2 •Sj \l 1 1 I" \1! 1 AC 1 r Note: B = buccal, Li = lingual, Lb = labial, M = mesial, and I = interproximal 239

C2-3-C Qiisidi'iinl loofh Presence Development Wenr Csiries Abscess ( iilculiis ; Ma\ill.ir\ Ki-ln M2 I \ell 1 - (l (i (l Ml I \ell 21 II 0 0 PI I \ell 5 II II II PI I \eM 1 Ii (I I) ( I \eI I I u II (I \lii\iliiirv I.ell C 2 Ac I-J 0 0 0 PI 2 \el-l 4 (I n II P2 2 \el-1 - 0 0 I) Ml 2 \ell 2? () 0 I M M2 2 Ael 1 l(> 0 0 0 Mi I \cl-l II (I (i (I Mandibular I ell \H 2 ,\el-l T 0 0 II M2 2 Ac 1-1 I'i II (I I I i M 2 \cl-l 2') I) 0 I I i P2 2 AeI I I II 0 I I i P. 2 \eM II I) li ( 2 \v I I II II I I i. I I h 12 2 \ell I) II 1 I i. I I b 11 <> \ \ M.iiklilnil.ii Ri'jlil I" ft \ \ 12 2 \, II ? II II I I i. I b (. 2 \v I I (I (I I I i PI 2 VII II II I I i P2 2 \cli 1 0 II 0 1 Ml 2 UN 2 ' II II I I i M2 2 Vj: il> 0 u I I l M3 2 Acl4 17 0 0 0 Note: B = buccal, Li = lingual, Lb = labial, M = mesial, and I = interproximal

C2-4-A Quadrant Tooth Presence Development Wear Caries Abscess Calculus Mandibular Left M3 2 Ael4 30 0 0 0 M2 2 Ac 14 12 0 0 0 Mandibular Right Ml 2 Acl4 25 0 0 IB Note: B = buccal, Li = lingual, Lb = labial, and I = interproximal

C2-4-B Quadrant Toolh Presence Development Wear Caries Abscess C iilculiis Mandibular Right M3 2 Acl4 4 0 0 0 Note: B = buccal, Li = lingual, Lb = labial, M = mesial, and I = interproximal 240

C2-4-C Quadrant looth Presence Development Wear Curies \hsccss Calculus M.indihiil.ii Kidil ( 1 VII (I u (i P! \cll i 0 0 0 P: 2 \J1 I 2 li li \li 2 VJI 22 I II 0 Note: B = buccal, Li = lingual, Lb = labial, M = mesial, and I = interproximal

C1-5-A Quadrant lonlh Presence Development Wear Caries thsress Calculus M.I\III.II\ Ki-hl Ml M: Ml 1ft 0 P2 0 PI \cl I • 1 h 12 \v.l4 II II \el-l I b M.IMII.II\ I ell II VI-I I I b 12 \cM i Mi ( Xv.14 I I b PI \cl1 1 M P2 UI 1 n Ml M: V.I4 Ml M.iiiihbiil.ii I ell Ml V I 1 12 I I I M2 Ml P2 PI ( ft n 0 I I i i: UI4 ft II (I II II M.iiidibiil.u Kmlii II 12 t PI P2 \LI I Ml M2 Mi Note: B = buccal, Li = lingual, Lb = labial, M = mesial, and I = interproximal 241

APPENDIX F: Results of Age and Sex Estimates Age Estimation Results

Individual Pubic Symphysis Auricular Epiphyseal Dental Dental Dental Final Age Suchey-Brooks Surface Union Eruption Formation Wear Estimation 2-.- \ ls-i'j "»1 . 18-21 18-21 3^-5ii 18-50 2-1-15 Ill-lft - - - Id 2-1-C I >2I . 25 18-21; 18-21 20-35 1 5-2-1 2-1-D -» -» - - "» ->. 2-1- \ - 18-21 18-21 15-50 - IX 2-4-li - 1S-21 ix-: i 2H-15 IS ?-•>-(• - 18-21 • 18-21 35-50 • IX 2-1-1) 2n - 2n 2--I-1 - Aduh 2-1-1 I - Vluh 2-1-1 2 - -\duli 2- l-l 1 - \ilull '2-1-1 I - \dull 2-1-1 - - Vluh 2-1-16 - \dull C2-4-F7 - Adult Bffc-5-A r26-70+ 40-44 iSPf*fuied 18-21t 18-21 50+ 40 *- A1-1-LM1 - Adult Al-l-LMT - Adilh A1-1-LM3 - Aduh - U-RM1 *w*n« Aduh U-RM2 - Adult U-RM3 - Adult Note = All ages are given in years. Sex Determination Results

,. ., . Traits of „ . .. Preauricular Nuchal Mastoid Supraorbital „. , „ Mental „ , Helmuth IndividuaT l „. . Sciatic „ . „ . _ ,„ Glabella _ . Head Sex Phenice - . . Sulcus Crest Process Margin Eminence _. (1985) NotcT h Diameter M M I'M M I'M 1 I'M M I'M

"2---1" M \1 M I'M M _ \1 \1 2-i-D I'l 1 I'l I'M •2-1- \ _ \1 - \1 PM :-4-n •?-i-c - IT PI :-4-D 1 IT Z-l-l. - 1 V C2-4-FltoC2- - U U 4-F7 I'l iff:: 'U' I'l I'l I'l 1 1 I'l A1-1-LM1 to - A1-1-LM3 u u ' "'' ".TJB^!™ ' ' *&&%%$. \" ' " ""Wv™* '4 A"" ' ';'-<-"'U- ?4 U-« SP- Note: F = female, PF = probable female, I = indeterminate sex, PM = probable male, M = male, and U = sex unknown. 244

APPENDIX G:

Summary of Oxygen Isotope Results 245

Summary of Oxygen Isotope Results

Individual Bone Bioapatite 518Q (%o) Tooth Bioapatite 518Q (%o) C2-1-A -4.7 -4 7 ('2-1-I1 -4/) - C2-M -4.') -1.7 C2-1-I) -l» - C2-4-A - -3.7 C2-4-H - -5.1 C2-4-C - -5.1 ( 2-4-1 1 -5.5 - ( 2-4-1 2 -4.4 - ( 2-4-1 ^ -4.1 - ("2-4-1 4 -4.(. - ( 2-4-1 5 -4.8 - C2-4-J o -4.3 - C2-4-I" -4.1 - (1-5- \ -4.7 .-4.4 \l-l-l Ml -V> - Al-I-I \I2 -3.5 - \l-l-l \H -1.8 - l-RMI -4.2 - 1 -RM2 -44 - I -1<\|1 -4.8 _ 246

APPENDIX H:

Results of Statistical Analyses 247

Test for Normality: Shapiro-Wilk

Variable Variable Division Statistic Degrees of Significance Freedom * I . - 1) l>2 20 P 0 10 • •1 \ - 1 W(l 2li P It 114 • " CVIII.IL:CII Vie d - 0 20 20 P 0 05' VV 1 " , ( - l«)l 2H P 0 1 1 vvl " .\ - I) l>? 20 P (1 08 I \ K.ilu i - i"-l 2n P il 2-1 J 0 I - )XS 18 P on2- l \ ( - 1 >^ IS P U4I Iii me Mii< ip.il lie t 1 - I4X ^ P ijy. l o1 ( . - IS8 "^ P n 2 » 1 UlHll 1 11.11 ill1' Biiuj\ilik' i 1 - )"1 ^ P 11.1)1'' Ne\ 0 C 1 cm.ilc MM 1 P 0 ;o o1 \ 1 cm.ilc ISI ^ P u22 '1 ime Pel •iod 1 o i 1 .II l> l" 1 .Me l 1.NS|I . l Id s P II Xf> I ale ( lassie i cp H P 0 21 l 1 nkii.'wn ID; % P II 2 > .1 \ 1 iiih K< 1 aid 1. iSli IX~ 5 P 0 2" 1 .111 ( l.lNMl »*> 1 H P II r 1 nkiioun )88 % P 0 11 i K ( 1 .iih !>• 1 .net 1.SM I Mil P 0 i ) 1 «ile t Liv.lv. M)2 11 P 0 10 1 Ilklli'WM I in ^ P US" 1 ~Z A \ , 1 .11 \\ Id 1 .Hi ( lc•>>l i 1 P OHO" 1 .III' ( l.lNNil. 1»M II P (l ^s I 1 Ilklh'Ull i8d 1 P li2S ,i ( 1 .III ( 1 ISNll i.SS 1 P il A Kurial'l.kp e 1 V |s| IWfi 1 P il "o \-l lumh 1 0 1 P II 86 ( -2- lumh m? W P II 12 1 "1 1 PI lib i«jii P 0 u. 1 Ilklli'WM i ni ^ P (I |<> <•> \ 11-.1 '.si I P 0 16 \-l lumh iv»2 P li 2'» l l -2- Ivimh ) ^ »J P (i fil Imiib i«>: i: P II 2S ^ 1 IlklK'VMl )SS i P 1" 11 i • > ( ( M i i~ i P iiX2 \-l lumh 19.S \ P 0 1" I -2- Imiib iSd ~ P 0 |'« Imiib iXS ID o i: ^ r 1 nkiiown 1 III * p US" \ t , I i>t )"2 1 p 11*2 \-l Iviiib 1 -\ ^ p II (III ( -2- 1 i'liih ).8'J - p 0 2~ 248

Test for Normality: Shapiro-•Wilk Continued Variable Variable Division Statistic Degrees of Significance Freedom

A" 11: .. coiiliiiucd Uunb 0S8 10 p-0.14 1 IlkllnWIl il S(> J) p = 0.28 (» '( . 1 cm ib O.'JX 3 p = 0.73 • Note: Isotopic values for Younger and Older Adults, Males and most isotopic values for Females could not be evaluated for normality because of small sample sizes (n < 2). *Approaches significance at the 0,05 level, so normal distribution of data cannot be assumed.. **Significant at the 0.05 level, so normal distribution of data cannot be assumed.

Spearman's Rank Order Correlation

rho N Significance d 'Ci..i n I 1 2(i p li-4 11 IliiLcn Yield (i 40 20 p o.ox- \\l '..I II IVJ 2n p (I (i»i ( \ -0 11 20 p (l Si) 1 •• I , -n (M IX p 0 >X \ ( -0 12 IS p- I) (O o"\ o (. 0 14 20 p o.^-l ( . Il.iivii N lekl (I I" 2") p () 4(> \\1 '»N 0.IX 2u p 0 Id i \ -II 02 2n p (1 v>S \ C, . -l) ,i» IX p (1II «>''( V C. . -O.iH IX p (I0| l U.-ne l I ii ()ii p li 2 » o'V.,, Tooth CI 0.68 p = 0.20 *Approaches significance at the 0.05 level. 249

Mann-Whitney U Test

Variable Z-value Significance Age s|St\.., -1.55 p- 0.12 -1.55 p-0.12 6'V,,, -1.55 p = 0.12 A Cin,,.^,! -1.55 p-0.12 Sex

5 'Ct0i -1.16 p -- 0.25 o']N -0.58 p - 0.56 8 'C|110 -0.78 p - 0.44

A ( I,K,.0,I -0.78 p - 0.44 -0.78 p •- 0.44 Sex (C2-3-I) Removed) s'V- -1.55 p -- 0.44 6 Ctl„ -0.78 p = 0.12 "I^v: o N 0.00* O MHO 0.00* A C I„„^„I Tombs -0.46 p - 0.64 5I5N -0.65 p -"- 0.52

S"C hui -1.03 p - 0.3 1 A"C hm-tol -0.80 p - 0.42 *Female sample size too small to perform Mann-U hitnc\ U Test (n = 1).

Kruskal-Wallis Test

Variable X Degrees of Freedom Significance '1 ime 1'eriii d i> I 2 (IX p 0 Mi i \ 2 81 P 0 12 * I! p II |d \ c, 0 02 p 1 011 linrfcil 1 > pe l il I. 2. I> p I) >J • ) \ 2 l«» p o s; i> ( 1 "(> P 0 23 \ ( 1 I "X p o (A il v. 1 20 p (I 2o Note: Preclassic and Simple were not included because each was represented by only a single case.