You are on page 1of 656

CAMBRIDGE STUDIES IN BIOLOGICAL AND EVOLUTIONARY ANTHROPOLOGY

CLARK SPENCER LARSEN

Bioarchaeology
Interpreting Behavior from the
Human Skeleton
Bioarchaeology
Interpreting Behavior from the Human Skeleton
SECOND EDITI ON

Now including numerous full color figures, this updated and revised edition of
Larsen’s classic text provides a comprehensive overview of the fundamentals of
bioarchaeology. Reflecting the enormous advances made in the field over the past
20 years, the author examines how this discipline has matured and evolved in
fundamental ways.
Jargon free and richly illustrated, the text is accompanied by copious case
studies and references to underscore the central role that human remains play in
the interpretation of life events and conditions of past and modem cultures, from
the origins and spread of infectious disease to the consequences of decisions made
by humans with regard to the kinds of foods produced, and their nutritional,
health, and behavioral outcomes. With local, regional, and global perspectives,
this up-to-date text provides a solid foundation for all those working in the field.

Clark Spencer Larsen is the Distinguished Professor of Social and Behavioral


Sciences and Chair of the Department of Anthropology at The Ohio State Uni­
versity in Columbus, Ohio. His research is focused primarily on biocultural
adaptation in the last 10 000 years of human evolution, with particular emphasis
on the history of health, well-being, and lifestyle. He collaborates internationally
in the study of ancient skeletons in order to track health changes since the late
Paleolithic. He is the author of 200 scientific articles and has authored or edited
30 books and monographs.
Cambridge Studies in Biological and Evolutionary Anthropology
Consulting editors
C. G. Nicholas Mascie-Taylor, University o f Cambridge
Robert A. Foley, University o f Cambridge

Series editors
Agustin Fuentes, University o f Notre Dame
Sir Peter Gluckman, The Liggins Institute, The University o f Auckland
Nina G. Jablonski, Pennsylvania State University
Clark Spencer Larsen, The Ohio State University
Michael P. Muehlenbein, Indiana University, Bloomington
Dennis H. O’Rourke, The University o f Utah
Karen B. Strier, University o f Wisconsin
David P. Watts, Yale University

Also available in the series


53. Technique and Application in Dental Anthropology Joel D. Irish ft Greg C. Nelson
(editors) 978 0 521 87061 0
54. Western Diseases: A n Evolutionary Perspective Tessa M. Pollard 978 0 521 61737 6
55. Spider Monkeys: The Biology, Behavior and Ecology o f the Genus Ateles Christina J.
Campbell 978 0 521 86750 4
56. Between Biology and Culture Holger Schutkowski (editor) 978 0 521 85936 3
57. Primate Parasite Ecology: The Dynamics and Study o f Host-Parasite Relationships
Michael A. Huffman ft Colin A. Chapman (editors) 978 0 521 87246 1
58. The Evolutionary Biology o f Human Body Fatness: Thrift and Control Jonathan C. K.
Wells 978 0 521 88420 4
59. Reproduction and Adaptation: Topics in Human Reproductive Ecology C. G. Nicholas
Mascie-Taylor ft Lyliane Rosetta (editors) 978 0 521 50963 3
60. Monkeys on the Edge: Ecology and Management o f Long-Tailed Macaques and their
Interface with Humans Michael D. Gumert, Agustin Fuentes, ft Lisa Jones-Engel
(editors) 978 0 521 76433 9
61. The Monkeys o f Stormy Mountain: 60 Years o f Primatological Research on the
Japanese Macaques o f Arashiyama Jean-Baptiste Leca, Michael A. Huffman, ft Paul
L. Vasey (editors) 978 0 521 76185 7
62. African Genesis: Perspectives on Hominin Evolution Sally C Reynolds ft Andrew
Gallagher (editors) 978 1 107 01995 9
63. Consanguinity in Context Alan H. Bittles 978 0 521 78186 2
64. Evolving Human Nutrition: Implications fo r Public Health Stanley UKjaszek. Neil
Mann, ft Sarah Elton (editors) 978 0 521 86916 4
65. Evolutionary Biology and Conservation ofTitis, Sakis and Uacaris Liza M. Veiga,
Adrian A. Barnett, Stephen F. Ferrari, ft Marilyn A. N'orconk (editors)
978 0 521 88158 6
66. Anthropological Perspectives on Tooth Morphology: Genetics. Eiobaion, Variation
G. Richard Scott ft Joel D. Irish (editors) 978 1 107 01145 8
67. Bioarchaeological and Forensic Perspectives on Violence: Flow Violent Death is
Interpreted from Skeletal Remains Debra L. Martin £t Cheryl P. Anderson (editors)
978 1 107 04544 6
68. The Foragers o f Point Hope: The Biology and Archaeology o f Humans on the Edge o f
the Alaskan Arctic Charles E. Hilton, Benjamin M. Auerbach, Et Libby W. Cowgill
978 1 107 02250 8
Bioarchaeology
Interpreting Behavior from
the Human Skeleton
SECOND EDITION

CLARK SPENCER LARSEN


The Ohio State University, USA

Cam bridge
U N IV E R SIT Y PRESS
Cambridge
U NIV ERSITY PRESS

University Printing House, Cambridge CB2 8BS, United Kingdom


One Liberty Plaza, 20th Floor, New York, NY 10006, USA
477 Williamstown Road, Port Melbourne, VIC 3207, Australia
314-321, 3rd Floor, Plot 3, Splendor Forum, Jasola District Centre, New Delhi - 110025, India
79 Anson Road, #06-04/06, Singapore 079906

Cambridge University Press is part of the University of Cambridge.


It furthers the University’s mission by disseminating knowledge in the pursuit of
education, learning and research at the highest international levels of excellence.
www.cambridge.org
Information on this title: www.cambridge.org/9780521547482
© Clark Spencer Larsen 2015
First edition © Cambridge University Press 1997
This publication is in copyright. Subject to statutory exception
and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.
First published 1997
Second edition 2015
A catalogue record for this publication is available from the British Library
Library of Congress Cataloging in Publication data
Larsen, Clark Spencer.
Bioarchaeology : interpreting behavior from the human skeleton / Clark
Spencer Larsen. - Second edition.
pages cm. - (Cambridge studies in biological and evolutionary anthropology)
ISBN 978-0-521-83869-6 (Hardback) - ISBN 978-0-521-54748-2 (Paperback)
1. Human remains (Archaeology) 2. Human skeleton-Analysis. I. Title.
CC77.B8L37 2015
930.1-dc23 2014031787
ISBN 978-0-521-83869-6 Hardback
ISBN 978-0-521-54748-2 Paperback
Additional resources for this publication at www.cambridge.org/Larsen
Cambridge University Press has no responsibility for the persistence or
accuracy of URLs for external or third-party internet websites referred to in
this publication, and does not guarantee that any content on such websites is,
or will remain, accurate or appropriate.
For Chris and Spencer and
In memory o f George J. Armelagos
(1936-2014), visionary scientist,
bioarchaeologist, friend, and mentor
C O NT E NT S

Preface to the Second Edition page xi


Preface to the First Edition xv

1 Introduction l

2 Stress and deprivation during growth anddevelopment andadulthood 7


2.1 Introduction 7
2.2 Measuring stress in human remains 8
2.3 Growth and development: skeletal 9
2.4 Growth and development: dental 25
2.5 Skeletal and dental pathological markers of deprivation 30
2.6 Adult stress 57
2.7 Summary and conclusions 64

3 Exposure to infectious pathogens 66


3.1 Introduction 66
3.2 Dental caries 67
3.3 Periodontal disease (periodontitis) and tooth loss 78
3.4 Nonspecific infection and disruption 86
3.5 Specific infectious diseases: treponematosis, tuberculosis,and leprosy 96
3.6 Specific infectious diseases: vectored infections 111
3.7 Summary and conclusions 112

4 Injury and violence 115


4.1 Introduction 115
4.2 Skeletal injury and lifestyle 116
4.3 Intentional injury and interpersonal violence 130
4.4 Medical care and surgical intervention 168
4.5 Interpreting skeletal trauma 172
4.6 Summary and conclusions 177

5 Activity patterns: 1. Articular degenerative conditions and


musculoskeletal modifications 178
5.1 Introduction 178
5.2 Articular joints and their function 179
5.3 Articular joint pathology: osteoarthritis 179
5.4 Nonpathological articular modifications 204
5.5 Nonarticular pathological conditions relating to activity 206
5.6 Summary and conclusions 212

6 Activity patterns: 2. Structural adaptation 214


6.1 Bone form, function, and behavioral inference 214
6.2 Cross-sectional geometry 215
x Contents
_ _______________________ __ ___________________________________________________ J
6.3 Histomorphometric biomechanical adaptation 246
6.4 Behavioral inference from external measurements 247
6.5 Summaiy and conclusions 255

7 Masticatory and nonmasticatory functions:


craniofacial adaptation to mechanical loading 256
7.1 Introduction 256
7.2 Cranial form and functional adaptation 256
7.3 Dental and alveolar changes 270
7.4 Dental wear and function 276
7.5 Summary and conclusions 300

8 Isotopic and elemental signatures of diet, nutrition, and lifehistory 301


8.1 Introduction 301
8.2 Isotopic analysis 302
8.3 Elemental analysis 347
8.4 Methodological issues in bioarchaeologicalchemistry 355
8.5 Summaiy and conclusions 356

9 Biological distance and historical dimensions ofskeletalvariation 357


9.1 Introduction 357
9.2 Classes of biodistance data 362
9.3 Biohistorical issues: temporal perspectives 368
9.4 Biohistorical issues: spatial perspectives 389
9.5 Summary and conclusions 401

10 Bioarchaeological paleodemography: interpreting


age-at-death structures 402
10.1 Introduction 402
10.2 Reconstructing and interpreting age-at-death profiles: it has been
mostly about mortality 404
10.3 Paleodemographers adopt the life table for age structure analysis 406
10.4 Addressing the assumptions of paleodemography 408
10.5 New solutions to interpreting age-at-death profiles in archaeological
skeletal series: it is really mostly about fertility not mortality 410
10.6 The elephant in the room: age estimates in archaeological skeletons 418
10.7 Summaiy and conclusions 419

11 Bioarchaeology: skeletons in context 422


11.1 Framing the contextual record 422
11.2 Framing the problems and questions: it is all about the hypothesis 424
11.3 Ethics in bioarchaeology 428
11.4 Bioarchaeology looking forward 429

References 433
Index 593

Color plates are to be found between pp. 320 and 321


PREFACE TO THE SECOND EDITION

It has been more than 15 years since the publication of the first edition of
Bioarchaeology: Interpreting Behavior from the Human Skeleton. The response
following its publication in 1997 was overwhelmingly positive – in reviews and
comments to me from virtually every corner of the globe. I credit Robert Benfer for
convincing me that a synthesis paper I wrote for Michael Schiffer’s book series,
Advances in Archaeological Method and Theory (Larsen, 1987), should be
expanded into a book-length treatment of the field. He made the case to me that
such a book would serve to define what bioarchaeologists do and give bioarch-
aeology a sense of identity and mission.
Since the publication of the first edition, I have been thrilled to see how the field
has matured and evolved, the increasing scientific rigor, the extraordinary volume
of work published, the high quality of the literature, the appeal that it has had for
new and upcoming generations of bioarchaeologists, the development of new
directions and advances, and the impressive increase in international and multi-
disciplinary collaborative research programs. With regard to new directions, we
have seen expansion in areas relating to links between the social and biological,
what some call “social bioarchaeology” (Agarwal & Glencross, 2011; Gowland &
Knüsel, 2006), and facets of it relating to identity, gender, and social and cultural
forces that leave their impression on the skeletal body (Knudson & Stojanowski,
2008, 2009; Larsen & Walker, 2010; Sofaer, 2006). In addition, there have been at
least two books published with Bioarchaeology as the primary title, one providing
a historical overview with reference to the United States (Buikstra & Beck, 2006)
and the other focusing on practice (Martin et al., 2013).
The advances in methods for the study of ancient skeletal and dental tissues
have expanded our understanding of past population health and lifestyle in ways
unfathomable or just on the horizon when the previous edition of the book was
published. As shown throughout the present volume, applications of the study of
ancient DNA to mobility and residence, disease diagnosis, and biology generally
are breathtaking (Kaestle, 2010). The advances made in genome-wide and sequen-
cing technology have given access to remarkable amounts of data, providing new
insights and perspectives on the human experience in the past. Similarly, imaging
technology has developed at a remarkable pace (Chhem & Brothwell, 2007;
Schultz, 2001). These advances have played a central role in the increasingly
interdisciplinary orientation of bioarchaeology (Armelagos, 2003; Zuckerman &
Armelagos, 2011). Fundamental to the development of bioarchaeology is its
comparative approach and its grounding in the scientific method and its approach
to discovery and problem solving. These strengths provide perspective on present
conditions, such as the human–environment interaction, evolution and adapta-
tion, and success and failure, and understanding of who we are today.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 20:01:11, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.001
xii Preface to the Second Edition

When I wrote the first edition, I had in mind a comprehensive volume, a


synthesis outlining what had been accomplished and future directions. So much
has been written since the first edition that this new edition does not attempt to
consider all advances that have appeared since the mid-1990s. Rather, I have
focused on key developments in areas that have more fully progressed in the last
couple of decades, as well as new and emerging areas, drawing on my own
experience and what has excited me most in bioarchaeological inquiry. In add-
ition, I provide a new stand-alone chapter on paleodemography. Demographic
structure of past populations provides insights into age profiles. More immediate
to bioarchaeology, age structure of archaeological skeletal series gives important
context for interpreting the variation seen in virtually all parameters discussed in
this book, ranging from diet and dietary change over the life course to reconstruc-
tion of lifestyle and activity via skeletal morphology and degenerative articular
pathology. I well understood the potential of paleodemography while I wrote the
first edition, but frankly, I thought that the area of study was in such disarray, that
I regarded a stand-alone chapter as preliminary and confusing. Since then,
however, there have been considerable advances made in paleodemography,
especially regarding the meaning of age structure for understanding population
dynamics and what is similar and different in comparing age structure of the dead
with vital statistics based on the living.
I also provide discussion of challenges that were presented in the concluding
chapter of the first edition, such as sample representation, the “osteological
paradox,” global perspectives, cultural patrimony, and the new world of genomics
and its importance to bioarchaeology and the study of the human past. Finally,
my own experience in bioarchaeology has widened greatly since I wrote the first
edition, especially resulting from the experience gained as codirector of two large
collaborative research projects, the Global History of Health Project and the
Çatalhöyük Bioarchaeology Project, and a field school in Medieval archaeology
and bioarchaeology (Field School Pozzeveri). Major funding from the US National
Science Foundation for the global project, the National Geographic Society and
the Templeton Foundation for the Çatalhöyük project, and the Italian government
for the field school and associated research program made all of this work
possible.
The preparation of the second edition of Bioarchaeology was an effort that
could have been completed only with a considerable amount of help. I received
advice on what the new edition should include or not include from Rimas
Jankauskas, Dale Hutchinson, Jackie Eng, Gwen Robbins Schug, Mike Pietru-
sewsky, George Milner, Sam Stout, Richard Scott, Graciela Cabana, Dan Temple,
George Armelagos, Tracy Betsinger, Maria Smith, Debbie Guatelli-Steinberg, Marc
Oxenham, Joel Irish, Marin Pilloud, Charlotte Roberts, Chris Stojanowski, and Kim
Williams. I owe a debt of gratitude to colleagues and students who read and
commented on individual chapter drafts. Thanks go especially to Helen Cho,
Giuseppe Vercellotti, Charlotte Roberts, Christina Torres-Rouff, Margaret Judd,
Pat Lambert, Tiffiny Tung, Michele Buzon, Bonnie Glencross, George Milner, Chris
Knüsel, Evan Garofalo, Chris Ruff, Libby Cowgill, Brigitte Holt, Marina Sardi, Rolo

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 20:01:11, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.001
Preface to the Second Edition xiii

González-José, Noreen von Cramon-Taubadel, Lesley Gregoricka, Sharon


DeWitte, Julia Giblin, Jess Pearson, Laurie Reitsema, Rob Cook, Annie Katzenberg,
Margaret Schoeninger, Christine White, Tracy Prowse, Mike Pietrusewsky, Chris
Stojanowski, Joel Irish, Marin Pilloud, Brian Hemphill, Leslie Williams, Ann
Stodder, Séb Villotte, and Britney Kyle. In addition, I benefited from advice from
Haagen Klaus, Dan Temple, Josh Sadvari, and Kathryn Marklein, who read the
entire manuscript and offered many substantive and helpful comments relating to
content and clarity.
I thank Tracey Sanderson, formerly of Cambridge University Press, for
approaching me to write the second edition, and to her successor, Martin Griffiths,
for sticking with me over the years of writing. Thanks also go to Ilaria Tassistro at
the Press for her assistance and skill as we moved the manuscript through the
production process and to Jeanette Mitchell for her excellent copy-editing.
I acknowledge the hard work by Sarah Martin and Kathryn Marklein in prepar-
ation of the bibliography.
I thank all of my friends and colleagues who provided photographs and other
figures. Those who are familiar with the first edition will note the considerable
expansion of the number of figures, to include many color images of pathological
conditions and other elements of morphology and biological variation. In add-
ition, I have increased the number of data and analysis graphs, largely in order to
help readers visualize research results discussed in the text. For their support in
providing photographs and graphs, thanks go especially to Chris Ruff, Haagen
Klaus, Kate Pechenkina, Tomasz Kozłowski, Valerie DeLeon, Sam Blatt, Megan
Brickley, Rachel Ives, Leslie Williams, Sam Scholes, Cory Maggiano, Pat Lambert,
Dale Hutchinson, George Milner, Charlotte Roberts, Jesper Boldsen, Eileen
Murphy, Kate Domett, Scott Haddow, Bonnie Glencross, Tim White, John Verano,
Tiffiny Tung, Margaret Schoeninger, Deborah Bolnick, Shannon Novak, Séb
Villotte, Chris Knüsel, Evan Garofalo, Jim Gosman, Richard Scott, Chris Schmidt,
Melissa Zolnierz, Lesley Gregoricka, Chris Stojanowski, and Joel Irish. Kathryn
Marklein provided considerable time and effort toward the development of the
electronic files of the more than 160 graphs, line drawings, and photographs.
I thank my parents, the late Leon Larsen and Patricia Loper Larsen, for introdu-
cing me at a very young age to old things and the past. I thank my undergraduate
professors at Kansas State University, especially my mentor and advisor, Patricia
O’Brien, and Professors William Bass and Michael Finnegan, and at the University
of Michigan, my PhD mentor and advisor, Milford Wolpoff, and Professors
Stanley Garn, Frank Livingstone, Loring Brace, David Carlson, Michael Zimmer-
man, and Roberto Frisancho for their inspiration and the training I received under
their collective direction. My fellowship stints at the Smithsonian Institution,
undergraduate and graduate, were strongly influential in the development of my
interests in bioarchaeology. I am especially grateful to Douglas Ubelaker, Dale
Stewart, Lawrence Angel, and Donald Ortner for their many stimulating discus-
sions, opportunities for research, and advice.
Since the publication of the first edition of Bioarchaeology, I moved to the
Department of Anthropology at The Ohio State University. At Ohio State, I have

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 20:01:11, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.001
xiv Preface to the Second Edition

been privileged to work with an extraordinary faculty and group of graduate


students, and to have access to superb research and teaching facilities. I am
grateful to the institution, my colleagues, and students for the stimulating intel-
lectual environment that helped to make this book possible.

Columbus, Ohio
May 1, 2014

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 20:01:11, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.001
PREFACE TO THE FIRST EDITION

The writing of this book was fostered by my involvement in a series of interdis-


ciplinary research programs undertaken in the southeastern (Florida and Georgia)
and western (Nevada) United States. I thank my collaborators, colleagues, and
friends who have been involved in this exciting research. With regard to field-
work, the following individuals and projects figured prominently in the develop-
ment of this book: David Hurst Thomas on St. Catherines Island, Georgia; Jerald
Milanich and Rebecca Saunders on Amelia Island, Florida; Bonnie McEwan at
Mission San Luis de Talimali in Tallahassee, Florida; and Robert Kelly in the
western Great Basin, Nevada. A number of individuals deserve special thanks
for their valuable contributions to the study of human remains from these regions:
Christopher Ruff, Margaret Schoeninger, Dale Hutchinson, Katherine Russell,
Scott Simpson, Anne Fresia, Nikolaas van der Merwe, Julia Lee-Thorp, Mark
Teaford, David Smith, Inui Choi, Mark Griffin, Katherine Moore, Dawn Harn,
Rebecca Shavit, Joanna Lambert, Susan Simmons, Leslie Sering, Hong Huynh,
Elizabeth Moore, and Elizabeth Monahan.
I thank the Edward John Noble Foundation, the St. Catherines Island Founda-
tion, Dr. and Mrs. George Dorion, the Center for Early Contact Period Studies
(University of Florida), the National Science Foundation (awards BNS-8406773,
BNS-8703849, BNS-8747309, SBR-9305391, SBR-9542559), and the National
Endowment for the Humanities (award RK-20111-94) for support of fieldwork
and follow-up analysis. Research leave given to me during the fall of 1991 while
I was on the faculty at Purdue University and a fellowship from Purdue’s Center
for Social and Behavioral Sciences during the spring and summer of 1992 gave me
a much needed breather from teaching and other obligations in order to get a
jump-start on writing this book. Preparation of the final manuscript was made
possible by generous funding from the University of North Carolina’s University
Research Council. I acknowledge the support – institutional and otherwise – of the
University of North Carolina’s Research Laboratories of Anthropology, Vincas
Steponaitis, Director.
A number of colleagues provided reprints or helped in tracking down key data
or literature sources. I especially thank John Anderson, Kirsten Anderson, Brenda
Baker, Pia Bennike, Sara Bon, Brian Burt, Steven Churchill, Trinette Constandse-
Westermann, Andrea Drusini, Henry Fricke, Stanley Garn, Alan Goodman, Gisela
Grupe, Donald Haggis, Diane Hawkey, Brian Hemphill, Frank Ivanhoe, Anne
Katzenberg, Lynn Kilgore, Patricia Lambert, Daniel Lieberman, John Lukacs,
Lourdes Márquez Morfín, Debra Martin, Christopher Meiklejohn, Jerome Melbye,
György Pálfi, Thomas Patterson, Carmen Pijoan, William Pollitzer, Charlotte
Roberts, Jerome Rose, Christopher Ruff, Richard Scott, Maria Smith, Dawnie
Steadman, Vincas Steponaitis, Erik Trinkaus, Christy Turner, Douglas Ubelaker,
John Verano, Phillip Walker, and Robert Walker.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:03:32, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.002
xvi Preface to the First Edition

Various versions of individual chapters and parts of chapters were read by


Kirsten Anderson, Brenda Baker, Patricia Bridges, James Burton, Stephen Church-
ill, Robert Corruccini, Marie Danforth, Leslie Eisenberg, Alan Goodman, Mark
Griffin, Gary Heathcote, Brian Hemphill, Simon Hillson, Dale Hutchinson, Anne
Katzenberg, Lyle Konigsberg, Patricia Lambert, Christine Larsen, George Milner,
Susan Pfeiffer, Mary Powell, Charlotte Roberts, Christopher Ruff, Shelley Saun-
ders, Margaret Schoeninger, Mark Spencer, Mark Teaford, and Christine White.
Ann Kakaliouras, Jerome Rose, and Phillip Walker generously donated their time
in the reading of and commenting on the entire manuscript. I am indebted to all of
the readers for their help in improving the clarity, organization, and content of
the book.
The organization of the bibliographic computer database was completed by
Elizabeth Monahan. Patrick Livingood helped in the preparation of figures. I thank
the following colleagues for providing photographs and figures: Stanley Ambrose,
Kirsten Anderson, David Barondess, Brian Hemphill, Charles Hildebolt, Dale
Hutchinson, George Milner, Mary Powell, Christopher Ruff, Richard Scott, Scott
Simpson, Holly Smith, Mark Teaford, Erik Trinkaus, Phillip Walker, and
Tim White.
A book like this is not written without a supportive press. I thank the Syndicate
of the Cambridge University Press and the Editorial Board of the Cambridge
Studies in Biological Anthropology – Robert Foley, Derek Roberts, C. G. N. Mascie-
Taylor, and especially, Gabriel Lasker – for their encouragement and comments,
especially when I proposed the idea of writing the book and what it should
contain. Most of all, I thank Tracey Sanderson, Commissioning Editor of Bio-
logical Sciences at the Press, for her help throughout the various stages, from
proposal to finished book.

Chapel Hill, North Carolina


August 28, 1996

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:03:32, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.002
1 Introduction

Many thousands of archaeological human skeletons are currently housed in various insti-
tutional repositories throughout the world. Some of these collections are extensive: The
Natural History Museum in London holds some 10 000 cataloged individuals, the Smithsonian
Institution has at least 30 000 skeletons, and the Bavarian State Collection more than
50 000 sets of human remains (Loring & Prokopec, 1994; McGlynn, personal communication;
Molleson, 2003). These and many other major collections around the world started during the
nineteenth century, mostly for purposes of collecting crania and other remains for investi-
gations of racial classification (Larsen & Williams, 2012; Little & Sussman, 2010) or as
“curiosities” without any manner of context (Taylor, 2014). While these motives for collection
have largely disappeared, these collections in Europe, North America, and elsewhere still
provide today an essential foundation for the study of past human variation and evolution.
The importance of the collections lies in the fund of biological information they offer for
interpreting lifeways of past peoples and for developing an informed understanding of the
history of the human condition generally.
One could argue that earlier generations of anthropologists may not have appreciated the
enormous value of human remains for interpreting the past, especially in view of the high
volume of excavated remains versus the remarkably low frequency of analysis and publica-
tion on those studied by biological anthropologists (Steele & Olive, 1989). Today, however,
human remains have become a key element of archaeological research and have contributed
to a burgeoning understanding of past population dynamics in general and the human
condition in particular. This is especially the case for developing and testing hypotheses and
drawing inferences about the human experience relating to diet, health, and lifestyle at all
levels, from the individual (various in Stodder & Palkovich, 2012; and see Knudson, Pestle
et al., 2012; Tiesler & Cucina, 2006) to large swaths of the globe (Steckel & Rose, 2002;
Verano & Ubelaker, 1992; White, 1999). In addition, the study of ancient human remains is
important in the ongoing discussions in anthropology about social organization, identity,
and the linkages between health and gender and health and status (various authors in
Gowland & Knüsel, 2006; Grauer & Stuart-Macadam, 1998; Knudson & Stojanowski,
2009). The study of these remains has played a fundamental role in investigating key
adaptive shifts in recent human evolution, such as the foraging-to-farming transition and
European exploration and colonization during the post-Columbian era (Bocquet-Appel &
Bar-Yosef, 2008; Cohen, 1989; Cohen & Armelagos, 1984; Cohen & Crane-Kramer, 2007;
Klaus, 2014a; Lambert, 2000a; Larsen, 2006; Larsen & Milner, 1994; Pinhasi & Stock, 2011),
and documentation and interpretation of specific disease histories (Palfi et al., 1999;
Powell & Cook, 2005; Roberts & Buikstra, 2003; Roberts et al., 2002). Especially in the last
decade or so, regional analysis has provided compelling insights into human adaptation in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:28:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.003
2 Introduction

broad perspective from a diverse range of settings globally (Domett, 2001; Fitzpatrick &
Ross, 2011; Hanson & Pietrusewsky, 1997; Hemphill & Larsen, 2010; Hutchinson, 2002,
2004, 2006; Judd, 2012; Lambert, 2000a; Larsen, 2001; Mushrif-Tripathy & Walimbe, 2006;
Ortner & Frohlich, 2008; Oxenham & Tayles, 2006; Pechenkina & Oxenham, 2012; Pietru-
sewsky & Douglas, 2002a,b; Robbins Schug, 2011b; Roberts & Cox, 2003; Ruff, 2010a;
Schepartz et al., 2009; Steckel & Rose, 2002; Stodder, 2008; Tung, 2012a; Weber et al., 2010;
Whittington & Reed, 1997; Williamson & Pfeiffer, 2003; Wright, 2006). These regional
studies have provided new perspectives on variation in human adaptation, including the
foraging-to-farming transition. For example, the experience of this transition in the western
hemisphere has shown a general decline in health, linked in part to maize farming. By
contrast, rice farming in Asia may have promoted better health, especially in relation to oral
health (Domett, 2001; Domett & Tayles, 2007; Douglas & Pietrusewsky, 2007; Oxenham,
2006). It is especially exciting to see the development of insights into past lives and lifestyles
in the regional context. In the mid-1980s, the record of poor health in the later prehistoric
American Southwest was beginning to emerge (Merbs & Miller, 1985). Building on this
record, discussions of conflict, its causes and consequences, and the crucial importance of
human remains in these discussions have developed, creating a forum for anthropologists
and other social scientists to engage in developing new solutions to long-standing problems.
In particular, it is through regional and continental comparisons that we are beginning to
understand the patterns and prevalence of violence and its effects on society, demography,
and health (Schulting & Fibiger, 2012).
Skeletal remains offer an important source of information for the study of human variation.
Archaeological skeletons from specific localities are more homogeneous both genetically and
in terms of the environments from where they came than are dissecting room or anatomical
skeletal series. Skeletons from the latter contexts are from many populations and highly
diverse circumstances. The use of archaeological series becomes especially important when
making conclusions about intra-population variability for a range of topics where sex and age
may be important factors.
Various surveys and manuals of human osteology and application to archaeological
settings are available (Baker et al., 2005; Bass, 2005; DiGangi & Moore, 2013; Roberts,
2009; Scheuer & Black, 2000, 2004; Schwartz, 2006; Ubelaker, 1999; White et al., 2012).
In order to address the incompatibility of different researchers’ methods and results, “stand-
ards” for skeletal data collection have been developed (Buikstra & Ubelaker, 1994). Although
dealing with the interpretive role in the study and documentation of human remains, these
works serve primarily as “how to” guides for bone identification and skeletal analysis and
not as resources for the investigation of broader issues in biological anthropology and sister
disciplines. The present book focuses on the relevance of skeletal remains to the study of the
human condition and human behavior generally; namely, how skeletal and dental remains
derived from archaeological settings reveal life history at both the individual and population
levels. It does not advocate a reliance on only human remains in order to tell the whole story
of the human condition. Rather, human remains represent a part of the broad sweep of data
derived from past settings. Human remains do not simply augment other data sources,
archaeological or historical. Rather, they provide perspectives and understandings of past
societies that pertain to human biology that simply may not be visible in other sources
(Perry, 2007).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:28:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.003
Introduction 3

The goal of this book is to provide a synthesis of bioarchaeology – the study of human
remains from archaeological contexts. Although the term was first applied to the study of
animal remains from archaeological settings (Clark, 1972), the focus of study then surfaced
with reference to the study of human remains in the regional “bioarcheological investigation”
of the lower Illinois River valley, an ambitious and innovative research program directed by
Jane Buikstra (1977a). This set into motion the future course of the field (Buikstra, 2006;
Knüsel, 2010; Larsen & Walker, 2010).
The field emphasizes integrative, interdisciplinary study. By doing so, the wide breadth of
bioarchaeology has engendered cross-fertilization between different disciplines, contributing
to its method and theory in approaching a wide diversity of problems. The field recognizes
the inextricable connection between biology and culture – one simply would not exist without
the other (Goodman, 2013). Just as human remains are a crucial component of study, it is the
context of these remains that provide us with meaning and substance. Context is a broad term
to include all potential sources, archaeological and otherwise, such as burial and social
inference, diet, climate, living conditions, and all else that is inferred or documented that
may inform our understanding of the people the skeletons represent.
The enormous potential of bioarchaeology for understanding the past is growing (Buikstra &
Beck, 2006; Katzenberg & Saunders, 2008; Larsen and Walker, 2010; Martin et al., 2013;
Stojanowski & Duncan, 2014). I believe that this is the case for two key reasons. First, in
contrast to earlier work that emphasized description, often poorly connected to any manner of
context, there is a growing interest in the central role that human remains play in understand-
ing patterns and trends in past societies. Second, the centrality of human remains for
understanding past biological, social, and behavioral dynamics has motivated an emphasis
on integrative research strategies, resulting in excavations of human remains having clear
agendas and questions that guide both their recovery and their study. Today, there is a very
different research profile, one that emphasizes the links between biological variation, health
and well-being, and behavior viewed broadly. This contrasts with the descriptive orientation
of earlier generations of osteologists, but it was the work of these predecessors that provided
many of the tools that served to form the field (Buikstra & Beck [2006], an historical treatment
of bioarchaeology).
This book takes a largely population perspective. However, individual-based case studies
are discussed, especially because collectively they help to build a picture of variability in
earlier societies. The population approach is critical for characterizing patterns of behavior,
lifestyle, disease, and other parameters that form the fabric of the human condition. The
discussion in the following pages also underscores the importance of culture and society in
interpreting population characteristics. Dietary behavior, for example, is highly influenced by
cultural and social norms of behavior. If an individual is taught that a specific food is “good”
to eat, then the consumption of that food item becomes fully appropriate in that cultural
context. Food is also an important indicator of a person’s place in society – access to resources
is influenced not only by where one lives but also by their identity and location within a
society. Hence, cultural and social factors play an essential role in determining diets of
individual members of a social group or society.
As I have made clear, the book is fundamentally driven by context and especially by the
placement of the biological record of skeletons in their archaeological context, recently called
“contextualization,” or the location of human remains within relevant archaeological and

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:28:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.003
4 Introduction

historical frameworks (Knudson & Stojanowski 2009; Thompson et al., 2014), and treatment
of the body in funerary settings (Duday, 2009). Contrary to the assertion that those who study
human remains have largely disengaged bones from context, whereby the physical anthro-
pologist “does not necessarily need the archaeologist once the archaeologist has excavated
bone” (Goldstein, 2006:376–377), there has been a growing movement focusing on the links
between archaeology and physical anthropology in the development of research programs
globally. Characterizing bioarchaeology as involving a disengagement of bodies from place
misrepresents the vast majority of research cultivated in this exciting discipline, especially
given the enormous progress of the field today.
Unlike many of the aforementioned guides to osteological analysis, this book is not a book
about methods. Certainly, methodological developments make possible much of the discus-
sion presented in the following chapters (see various authors in Katzenberg & Saunders,
2008). Moreover, technological advances, such as imaging, have provided new insights
unimaginable a decade ago (various in Chhem & Rühli, 2004). However, this book focuses
on how human remains inform our understanding of the past. By doing so, this book is
intended to feature the various insights gained about human behavior and biology rather than
to describe or evaluate specific methods and techniques of skeletal analysis. This approach is
central to the biocultural perspective offered by anthropologists – we must seek to envision
past populations as though they were alive today, and then ask what information drawn from
the study of skeletal tissues would provide understanding of them as functioning, living
human beings and members of populations. Nor is this book a critical review, highlighting the
shortcomings of the field or what bioarchaeologists should be doing, but are not.
Bioarchaeological findings are important in a number of areas of scientific and scholarly
discourse. Within anthropology, the use of human remains in interpreting social behavior
from mortuary contexts is especially fruitful (Artelius & Svanberg, 2005; Beck, 1995; Buik-
stra, 1977a; Chapman et al., 1981; Chesson, 2001; Gowland & Knüsel, 2006; Humphreys &
King, 1981; Knudson & Stojanowski, 2009; O’Shea, 1984; Parker Pearson, 2000). The story
human remains tell is also reaching an audience outside anthropology. There is an increase in
use of bioarchaeological data in history, economics, and nutrition science. Skeletal studies of
nutrition, disease, and related topics, and the importance of human remains in developing a
broader understanding of economic history are opening a new path of research interest
(Steckel & Rose, 2002). The emerging role of skeletal remains in the study of the human
condition was underscored by the eminent historian, John Coatsworth (1996:1), who high-
lights the “masses of evidence” provided from bioarchaeological investigations and the
important role they play in understanding historical developments.
Breakthroughs have been made in the analysis of various body tissues in archaeological
settings, including hair, muscle, skin, and other soft tissues (Arriaza, 1995; Asingh &
Lynnerup, 2007; Aufderheide, 2003; Brothwell, 1987; Chamberlain & Parker Pearson, 2001;
Cockburn et al., 1998; Hansen et al., 1991; Lynnerup et al., 2002; Stead et al., 1986). The
discussions presented in this book focus largely on skeletal and dental tissues. Building on the
study of human remains, the unifying theme in this book is behavioral inference. My
discussion of behavior is not limited to physical activity; rather, it is considered in a wider
perspective, including (in order of appearance in the book) physiological stress, exposure to
pathogenic agents, injury and violence, physical activity, masticatory and extramasticatory
uses of the face and jaws, dietary reconstruction and nutritional inference, population history,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:28:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.003
Introduction 5

and social factors and how they influence health and lifestyle. Simply, almost everything
available in the study of human remains has behavioral meaning.
Bioarchaeology is represented throughout the world. This book draws upon a sample of this
record in illustrating important points and issues. The book deals with all regions of the globe,
but especially in those areas that have been studied in depth by bioarchaeologists. One of the
exciting advances in the field in the last decade or so is the proliferation of bioarchaeological
treatments outside North America, especially Europe, Asia, and the Pacific (Domett, 2001;
Lynnerup, 1998; Murphy, 2002, 2003; Ortner & Frohlich, 2008; Oxenham & Tayles, 2006;
Papathanasiou, 2001; Pechenkina & Oxenham, 2013; Pietrusewsky & Douglas, 2002b; Rife,
2012; Roberts & Cox, 2003; Ruff, 2010a; Schepartz et al., 2009; Whittington & Reed, 1997;
and many others). Reflecting this increasingly international bioarchaeology are the inter-
national, multidisciplinary collaborations around the world, including in Latin America,
Europe, Asia, and Africa.
Various points made in the book are addressed by contrasting and comparing data sets
from skeletal assemblages representing human populations from different levels of socio-
political complexity and differing subsistence regimes. Because of the vagaries of dietary
reconstruction in the archaeological past, anthropologists usually characterize human groups
broadly, using terms such as “foragers” or “farmers.” The reader should recognize that these
terms are often overly simplistic and do not convey the underlying complexity of human
adaptive systems adequately. Nevertheless, these categories help us to understand broad
behavioral and adaptive features of different groups better, and therefore, are the starting
points to facilitate yet more specific and contextual reconstructions and interpretations of
past lifeways. Of far more importance to the focus of this book is that these contrasts and
comparisons add an important dimension to the growing discussion in anthropology oriented
toward the understanding of the causes and consequences of adaptive and behavioral shifts in
the past and present.
A fundamental barrier to understanding health in today’s populations is the very narrow
temporal window in which they have been studied. The prevalence records for osteoarthritis
and oral degenerative conditions, for example, are limited largely to the last several decades
when large-scale epidemiological studies of living populations were first undertaken (Arden &
Nevitt, 2006; Blinkhorn & Davies, 1996; Issa & Sharma, 2006; Jordan et al., 2007; Pilot, 1998;
Sreebny, 1982; Woodward & Walker, 1994), representing approximately only the last 0.1% of
the history and evolution of our species. Biomedical, experimental, molecular, and behavioral
studies of skeletal and oral degenerative conditions have been investigated in great detail in
human populations for this 0.1% window of time. However, these studies are limited in scope,
focusing primarily on remarkably gross categories of human variation. They are deficient
because they underrepresent and mischaracterize human biological variation, reducing the
variation to simple dichotomous (or other limited) comparisons having little to do with real
biological or social variation. Analysis of the recent biomedical and epidemiological literature
characterizes variation according to “race” (e.g., White vs. Black vs. Hispanic), geography
(mostly United States and western Europe), sex (men vs. women), diet (e.g., access to refined
sugar), and socioeconomic status based on income level (upper-, middle-, and lower-income),
in addition to the very narrow temporal window (Allen, 2010; Dominick & Baker, 2004;
Jordan et al., 2007; Pilot, 1998). The (mis)characterization of human variation has important
implications for understanding an increasingly diverse society in the United States, far more

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:28:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.003
6 Introduction

diverse than the racial categories so prevalent in biomedical research would seem to imply. By
expanding understanding of diversity, in terms of both broad temporal and geographic
perspectives, bioarchaeology as an anthropological subdiscipline provides a more informed
understanding of health in the present through a consideration of heath in the past.
Anthropologists, economic historians, and other social scientists have long recognized that
humans in the past and present are extraordinarily diverse in their food consumption
practices, social habits, workloads, and other behavioral characteristics that collectively
characterize health and lifestyle. The biomedical and epidemiological literature on degenera-
tive conditions is known from a limited, if not insufficient record. Moreover, this record is
limited to the study of populations that have mitigated some of the conditions that influence
the prevalence of degenerative conditions most strongly, namely industrialized societies
having access to healthcare, adequate nutrition, and labor-saving technology. For example,
prevalence of dental caries in many developed countries has been reduced due to the
introduction of fluoride in drinking water, which is certainly the case in the United States
(McDonagh et al., 2000). In addition, owing to work-saving technology, workload has greatly
declined in developed countries especially. If physical activity is the primary factor in
determining prevalence and pattern of osteoarthritis, for example, then one could predict a
decline in its prevalence, especially in recent societies. That is, as advances in technology
essentially replace what the body used to do in work and activity generally, we should expect
to see long-term trends in terms of a decline in osteoarthritis. This hypothesis has not been
tested using the kind of systematic approach offered by bioarchaeology.
Bioarchaeology offers a unique opportunity to study a much more diverse sampling of
humanity, namely the last 10 000 years of more than seven million years of human evolution,
greatly extending the time framework for characterizing the biocultural context surrounding
some of the most important changes and developments in the evolutionary history of our
species. Arguably, it is this small percentage of our evolution where crucial developments
occurred that set the stage for the rise of modern civilization, namely farming, appearance of
complex societies, urbanization, and industrialization.
Many studies of human remains from archaeological contexts focus on a single population
without actively linking the analysis of these remains to the context (climate, diet, time,
culture, settlement, and economic system) from which they derive. A central aim of bioarch-
aeology is to establish a comprehensive record of skeletal and dental conditions in relation to
prevalence and pattern to develop an understanding of behavior and the costs and conse-
quences of particular lifestyle circumstances and conditions. Bioarchaeology provides an
unmatched record of health and life conditions in the past, thereby extending our understand-
ing of diversity in geography, cultures, and time that is simply not possible with sole reliance
on limited health attributes of living populations.
Human skeletal and dental tissues are remarkable records of lives led in the past, what
Stanley M. Garn referred to as “a rich storehouse of individual historical events” (1976:454).
This book provides a tour of the vast holdings in this storehouse, displaying the knowledge
gained about earlier peoples based on the study of their mortal remains.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:28:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.003
Stress and deprivation during growth
2 and development and adulthood

2.1 Introduction
Physiological disruption resulting from impoverished environmental circumstances – “stress” –
is central to the study of health and well-being and the reconstruction and understanding of
health, adaptation, and behavior in both earlier and contemporary human societies. Stress
involves disruption of homeostasis, or the maintenance of a constancy of conditions that keep
the body’s internal environment stable. Stress is a product of three key factors, including
(1) environmental constraints, (2) cultural systems, and (3) host resistance. Goodman and others
(Goodman, 1991; Goodman & Armelagos, 1989; Goodman & Martin, 2002; Goodman, Martin
et al., 1984, 1988; Klaus, 2012; Martin et al., 1991) modeled the interaction of these factors at
both the individual and population levels (Figure 2.1). This model of health and context
emphasizes the role of environment in providing both the resources necessary for survival and
the stressors that may affect the health and well-being of a population, yet includes the profound
influence of culture in health outcomes. Cultural systems serve as protective buffers from the
environment, such as shelter and clothing as buffers against temperature extremes. Cultural
systems can be highly effective at mitigating behaviors necessary for extraction of important
nutrients and resources from the environment, thus supporting the ability to maintain stability.
It appears impossible for the full spectrum of stressors in an environment to be buffered against;
some inevitably slip through the filters of the cultural system. In these instances, the individual
may exhibit a biological stress response observable at the tissue level (bones and teeth).
Importantly, stress, buffering systems, and tissue-level responses are not linked by a simplistic,
linear relationship. Instead, they can interact with and influence other variables within other
levels of the model. Physiological disruption feeds directly back into environmental constraints
and cultural systems. This model makes clear that health is a key variable in the adaptive process.
Biological stress has significant functional consequences for the individual and for the
society in which they are living. Elevated stress and associated disruption of homeostasis may
lead to a state of functional impairment, resulting in diminished cognitive development and
work capacity. The reduction in work capacity can be detrimental if it impedes the acquisition
of essential resources (e.g., dietary) for the maintenance of the individual, the population, and
the society. If individuals of reproductive age are affected by poor health, then decreased
fertility may be the outcome. Ultimately, the success or failure of a population and its
individual constituents to mitigate stress has far-reaching implications for behavior and the
functioning of society.
Stress and developmental instability in early life, prenatal and postnatal, have important
implications for health outcomes in later life. David Barker’s research on chronic diseases in
middle age reveals that individuals with low birth weight and nutritional deprivation in early

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
8 Stress and deprivation during growth and development and adulthood

Environmental Constraints Culturally Induced Stressors


Limited Resource Social hierarchy
Ecological Variables Differential Access to Resources
Physiogeography Sociopolitical Structure

Cultural Buffering Systems


Technological and Behavioral Adaptations

Genome, Developmental Pathways, and Epigenetic Inputs

Host Resistance
Feedback
Loop
Physiological Disturbance
Growth Disruption
Feedback Disease
Loop Death

Poplulation-Level Outcomes
Increased Morbidity: Epigenetic Effects
Decreased Work and Reproductive Capacity
Sociocultural Disruption and Instability

Biological Adjustment Behavioral


and Adaptation Alteration

Figure 2.1 Biocultural model for interpreting stress. This model emphasizes the primacy
of environmental constraints and cultural influences on outcomes in health and well-being.
(Adapted from Klaus, 2012; reproduced with permission of author and University Press
of Florida.)

life are more predisposed to earlier death and greater frequency of chronic disease, including
cardiovascular disease, hypertension, and type 2 diabetes, than are individuals with normal
birth weight and sufficient prenatal and early natal nutrition (Barker, 2001, 2012; Barker
et al., 2002; Beltrán-Sánchez et al., 2012; Hales et al., 1991; Harding, 2001; and see Jo-
vanovic, 2000; various in Henry & Ulijaszek, 1996; and others). Similarly, low birth weight
infants show lower bone mass as adults than infants with normal birth weight (Antoniades
et al., 2003; Cooper et al., 1995; Gluckman et al., 2008). Experimental evidence in laboratory
animals suggests that poor prenatal environments are also a risk factor for growth stunting
over the life course in general (Dancause et al., 2012). While the prenatal environment does
not determine health consequences in adulthood, it appears to have a profound role to play in
programing circumstances in later life that promote poor health. A growing record supports
Barker’s developmental origins hypothesis regarding the profound influence of intrauterine
stresses in early life on later health, morbidity, and mortality.

2.2 Measuring stress in human remains


Biological anthropologists employ a variety of skeletal and dental stress indicators that can be
measured empirically. Use of multiple indicators gives a comprehensive understanding of
stress and adaptation in past populations (Goodman & Martin, 2002). The multiple-indicator

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.3 Growth and development: skeletal 9

approach stems from the recognition that health is a composite of nutrition, disease, and other
aspects of life history. Simply, simultaneous study of multiple, but independent, markers of
stress provides a more reliable and informed understanding of composite health as docu-
mented in a skeletal series. Contrary to medical models of health, stress and disease (see
Chapter 3) represent a continuum rather than a presence versus absence phenomenon, with
respect to both the population and the individuals that comprise it.

2.3 Growth and development: skeletal


2.3.1 Growth rate
Rate or velocity of growth in humans shows considerable variation in the comparison of
different human populations, present and past (Bogin, 1999; Hoppa & FitzGerald, 1999; Ruff
et al., 2013; Saunders, 2008). Nevertheless, human growth is punctuated by two intensive
periods of activity from birth through adolescence. The first period pertains to a great increase
in growth velocity during infancy, falling off soon after the first two years of life. The second
involves another marked increase during adolescence, then declines and reduces to zero
growth when epiphyseal fusion of the long bones (femur, tibia, fibula, humerus, radius, and
ulna) and other skeletal elements is complete, marking full skeletal maturity (Crews & Bogin,
2010). Growth rate is widely recognized as a highly sensitive indicator of health and well-
being of a community or population (Bogin, 1999; Eveleth & Tanner, 1990; Foster et al., 2005;
Lewis, 2007).
Various factors affect growth, such as genetic influences, growth hormonal deficiencies,
and psychological stress (Bogin, 1999; Eveleth & Tanner, 1990; Gray & Wolfe, 1996; Ruff
et al., 2013), but the preponderance of evidence underscores stressors that are produced by
adverse environments, especially poor nutrition, in shaping growth and development
(Foster et al., 2005; Leonard et al., 2000; Moffat & Galloway, 2007). Infectious disease
can also contribute to poor growth, such as episodic diarrheal disease and circumstances
involving poor sanitation and suboptimal living conditions that ultimately reduce nutrition
at the cellular level (Cardoso, 2007; Jenkins, 1982; Martorell et al., 1977; Moffat 2003).
Nutrition and disease have a synergistic relationship whereby poorly nourished juveniles
are more susceptible to infection, while infectious disease reduces the ability of the body to
absorb essential nutrients (Keusch & Farthing, 1986; Scrimshaw, 2003, 2010; Scrimshaw
et al., 1968).
Children raised in impoverished environments in developing nations or in stressed settings
in developed nations generally are small in size for their age (Bhargava, 1999; Bogin, 1999;
Crooks, 1999; Eveleth & Tanner, 1990; Foster et al., 2005). Among the best-documented
populations are the Mayan Indians of Mesoamerica, who show retarded growth in comparison
with reference populations from less adverse settings (Crooks, 1994). In Guatemala City,
Guatemala, well-fed upper-class children are taller than poorly nourished lower-class children
(Bogin & MacVean, 1978, 1981, 1983; Johnston et al., 1975, 1976). Additionally, unlike the
markedly slower growth in lower-class children, upper-class children have comparable
growth to Europeans. The cumulative differences between Mayan and European children
are especially pronounced for the period preceding adolescence, suggesting that growth
during the early years of childhood is the most sensitive to environmental disruption in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
10 Stress and deprivation during growth and development and adulthood

comparison with other life periods (Bogin, 1999). During adolescence, genetic influence on
growth is more strongly expressed than in childhood (Bogin, 1999; Bogin & Loucky, 1997).
Juveniles have been growing taller over much of the twentieth century in industrialized
countries and in some developing nations. This secular trend in growth is related to a variety
of environmental and cultural changes, including improvement in food availability and
nutrition, sanitation, reduction of infectious disease, and increased access to Western health-
care. As the environment improves, growth increases. On the other hand, declines in growth
velocity are well documented, especially during periods of dietary deprivation in wartime
settings, famines, and economic crises (Eveleth & Tanner, 1990; Himes, 1979; Komlos 1994).
This link between growth status and environment is well documented via analysis of histor-
ical data. Comparisons of heights of British school children from various regions and eco-
nomic circumstances for the period of 1908 to 1950 show that children were generally shorter
in areas experiencing high unemployment (e.g., Glasgow, Scotland) than in other regions with
more robust economies (Harris, 1994). These differences were especially pronounced during
the severe economic depression in the late 1920s when nutritional and general health of
children of unemployed parents declined. Similarly, growth velocity and attainment per age
increased in post-World War II following the amelioration of negative socioeconomic condi-
tions (Cardoso & Gomes, 2009; Malina et al., 2010; Tanner et al., 1982). In post-1945 Poland,
relatively greater increases in growth were documented in higher socioeconomic groups
(Bielicki & Welon, 1982).
Beginning with Francis Johnston’s pioneering work on childhood growth based on the
study of archaeological skeletons (Johnston, 1962, 1969), a range of studies have extended
our understanding of growth rates in past societies. That is, the general pattern of juvenile
growth in archaeological populations is broadly similar to that in living populations (Arme-
lagos et al., 1972; Boldsen, 1995; Edynak, 1976; Hillson et al., 2013; Hoppa, 1992; Huss-
Ashmore, 1981; Johnston, 1962; Lewis, 2002; Mays, 1999; Merchant & Ubelaker, 1977;
Molleson, 1995; Ribot & Roberts, 1996; Ruff et al., 2013; Ryan, 1976; Saunders, 2008; Sciulli &
Oberly, 2002; Storey, 1992a, 1992b; Sundick, 1978; Walimbe & Gambhir, 1994; Walker, 1969;
and see later). The congruence of growth in past and living groups suggests that there has not
been a change in the general pattern of growth in recent human evolution (Saunders, 2008).
That is, patterns and processes of growth in known human populations appear to have been
present for at least the last 10 000 years of our evolutionary history, and likely longer.
Some populations appear shorter for age than others. Analysis of juvenile long bones from
prehistoric North America reveals evidence of growth retardation in agricultural and mixed
subsistence economies versus foragers. In children younger than six years of age in the
prehistoric lower Illinois River valley, matching of femur length to dental age reveals growth
suppression in late prehistoric (Late Woodland period) maize agriculturalists in comparison
with earlier foragers (Middle Woodland period) (Cook, 1979, 1984). Cook (1984) concluded
that the decline in growth was due to a decrease in nutritional status with the shift to a
protein-poor maize diet. Children short for their age during the later prehistoric period tend to
express a higher frequency of stress indicators (e.g., porotic hyperostosis, enamel defects) than
children who are tall for their age, lending further support for nutritional deficiency as a prime
factor contributing to growth retardation.
Lallo (1973; and see Goodman, Lallo et al., 1984) found, in addition, a decrease in the
growth of femur, tibia, and humerus diaphysis lengths and circumferences during the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.3 Growth and development: skeletal 11

Mississippian period (AD 1200–1300) in comparison with earlier periods (AD 950–1200) in the
central Illinois River valley. Dietary change during this time involved a shift from mixed
foraging and farming to intensive maize agriculture. Growth during the period between two
and five years of age was especially slow, which Goodman and coworkers (1984) conclude
reflects an increase in physiological stress due to poorer nutrition and the presence of other
stressors during the later prehistoric occupation of the region.
The impact of increased stress loads due to the combined effects of undernutrition,
European-introduced infectious disease (e.g., smallpox, measles), warfare, and increased
social disruption has been investigated in the late prehistoric and contact-era Arikara
Indians of the upper Missouri River valley (Jantz & Owsley, 1984a, 1984b, 1994a; Owsley &
Jantz, 1985). Matching of long bone lengths (femur, tibia, humerus, radius) to dental age in
perinatal (late fetal/early neonatal) and other juvenile skeletons reveals that late postcontact
era (AD 1760–1835) Arikara juveniles are smaller than early postcontact (AD 1600–1733)
juveniles, suggesting declining health status as European influence and encroachment into
the region by other tribes increased. Comparison of postcontact Arikara with healthy upper
middle-class Euroamericans from Denver, Colorado confirms a pattern of slower growth in
early childhood in the Native American group (Hillson et al., 2013; Ruff et al., 2013)
(Figure 2.2). In contrast, growth in young children from Neolithic Çatalhöyük, Turkey is
strongly similar to the Denver growth study sample, indicating relatively better nutrition
and health overall, a pattern consistent with other skeletal indicators (Hillson et al., 2013;
Ruff et al., 2013) (Figure 2.2).

200

180

160

140
Stature (cm)

120

100

80

60

0 2 4 6 8 10 12 14 16 18 20 22 24
Age (years)
Figure 2.2 Fitted curves for estimated juvenile statures for protohistoric Arikara (open
circles and dashed lines), Neolithic Çatalhöyük (solid diamonds and solid line), and modern
Denver population (asterisks and dotted line). (From Ruff et al., 2013; © 2012 Wiley
Periodicals, Inc.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
12 Stress and deprivation during growth and development and adulthood

The interaction between stress and population mobility has been examined in the compari-
son of Late Archaic-period foragers from the Carlston Annis Bt-5 site, Kentucky (3992–2655
BC) and Late Woodland foragers from the Libben site, Ohio (AD 800–1100) (Mensforth, 1985).
Archaeological evidence indicates that Late Archaic populations were highly mobile and
exclusively dependent on wild plants and animals. In contrast, maize was consumed by the
Libben population, but it was of minor dietary significance. For both groups, nutrition appears
to have been adequate (Mensforth, 1985). Comparisons of tibia lengths reveal a general
similarity between the two groups from birth to six months and from four years to 10 years.
For juveniles aged six months to four years, Libben tibiae are shorter than Bt-5 tibiae. The
growth period between six months and four years – the period differing most between Bt-5
and Libben populations – is highly sensitive to metabolic disruption. During this period, an
infant undergoes weaning, involving the shift from breast milk, a relatively stable, nutritious
food source, to exclusive solid food, often less stable in availability, less digestible, and less
nutritious. Passive immunities derived from consumption of breast milk are lost during
weaning during this period of life (Popkin et al., 1986). These immunities are crucial for early
health and well-being because the child’s immune system is not fully developed until after
five years of age (Newman, 1995). Consistent with this outcome of compromised immunity,
Mensforth (1985) found a high prevalence of nonspecific periosteal infections in the Libben
infants, suggesting that high levels of infectious disease in infancy and young childhood
contributed to growth retardation. Thus, in comparison with the Bt-5 population, the Libben
population experienced the effects of greater sedentism and community size that fostered
poor sanitation, elevated infectious disease, and poor health.
Comparison of archaeological samples with a modern reference population (Denver, Color-
ado; Maresh, 1970) confirms the presence of growth rate suppression in children from
archaeological settings. That is, calculation of the average percentage of adult size attained
in successive age groups for the major long bones reveals a slow-down in most settings
(Humphrey, 2003), consistent with the general finding that children in archaeological con-
texts are shorter than their modern counterparts (Lewis, 2007; Mays, 1999). Lovejoy and
coworkers (1990) argue that widespread infection was the chief cause of growth retardation in
the Libben setting. They suggest that inflammation would result in an increased production of
cortisol, the major natural glucocorticoid, which results in limitation of growth and availabil-
ity of amino acids. Thus, elevation of infection in the Libben population may have had a
strong influence on growth generally (Lovejoy et al., 1990).
Historic-era skeletal series furnish important insights into stress in the recent past. Saunders
and coworkers (1995, 2002; Humphrey, 2003) analyzed growth data available from a large
series of juvenile remains from the St. Thomas’ Anglican Church cemetery in Belleville,
Ontario. The cemetery was used by a predominantly British-descent population during the
period of 1821 to 1874. Comparisons of femur length from juveniles buried in the cemetery
with a tenth-century Anglo-Saxon series from Raunds, England, and modern growth data
from Denver, Colorado (Maresh, 1970), indicate a strong similarity in overall pattern of
growth between the three groups (Figure 2.3). The two cemetery samples are temporally
separate, but share general ethnic origins with the modern US population. Figure 2.3 shows
that the St. Thomas’ series is slightly shorter than the modern series for age. The Raunds series
is considerably shorter than either of the other groups, which is to be expected given the
inferior living standards of tenth-century England. With regard to the St. Thomas’ skeletons,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.3 Growth and development: skeletal 13

400

350
Diaphyseal length (mm)

300

250

200

150

100

50
0 1 2 3 4 5 6 7 8 9 10 11 12 13
Age (years)
Figure 2.3 Fitted curves for femoral diaphyseal length for the nineteenth-century
St. Thomas’ Church cemetery (dotted line), tenth-century Raunds Anglo-Saxon skeletons
(dashed line), and twentieth-century Denver, Colorado, living population (solid line).
(From Saunders & Hoppa, 1993; reproduced with permission of John Wiley & Sons, Inc.)

Saunders and coworkers suggest that juveniles died from acute causes and not chronic
conditions (e.g., chronic infections or chronic undernutrition) that would result in a decrease
in skeletal growth. Children younger than two years of age had slightly lower growth rates
than those of modern twentieth-century populations. They regard this as perhaps representing
stresses associated with poor maternal health and prenatal growth.
Similarly, on the north coast of Peru, comparison between late pre-Hispanic- and
postcontact-period femoral growth velocity reveals markedly slower growth among the
indigenous children, especially at ages 5, 10, and 12 years in the colonial Lambayeque Valley
(Klaus & Tam, 2009). This decrease in the rate of growth takes place in a setting of introduced
European disease, increased prevalence of other skeletal stress markers, poor diets, and
sociopolitical marginalization.
Analysis of juvenile cortical bone growth via measurement of cortical thickness provides a
complementary source of information to long bone lengths. In living populations, deficiencies
in cortical bone mass are present in groups experiencing undernutrition (Agarwal, 2008;
Agarwal & Glencross, 2011; Cooper et al., 1995; Frisancho, Garn et al., 1970; Garn, 1970;
Garn et al., 1964; Gluckman et al., 2008; Himes, 1978; Himes et al., 1975). The pioneering
investigation by Garn and coworkers (1964), for example, showed that malnourished Guate-
malan children have reduced cortical bone in comparison with well-nourished reference
groups. Although bone lengths increased during periods of growth recovery, cortical thickness
continued to show deficiencies due to earlier episodes of bone loss. Thus, growth recovery

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
14 Stress and deprivation during growth and development and adulthood

may involve an increase in bone length (and attained height), but not bone mass (Antoniades
et al., 2003; Huss-Ashmore, 1981; Huss-Ashmore et al., 1982).
Cortical bone mass is a sensitive indicator of environmental disturbance in archaeological
settings. Comparison of femoral cortical thickness from Middle Woodland (Gibson site) and
from Late Woodland (Ledders site) series from west-central Illinois reveals a reduction in bone
mass in young children (24–36 months), the presumed time of weaning and increased dietary
stress (Cook, 1979). In contrast, Hummert and coworkers (Hummert, 1983; Hummert & Van
Gerven, 1983; Van Gerven et al., 1985) documented increased cortical bone deficiencies in
exclusively older children from the early to late Christian periods in Sudanese Nubia (c. AD
550–1450). Long bone lengths of Nubians are shorter in the early Christian period than in the
late Christian period, which may be due to nutritional deficiencies and bacterial and parasitic
infections (Hummert, 1983; Hummert & Van Gerven, 1983). Increasing political autonomy
during the later Christian period may have served to improve living conditions, resulting in
better growth status and health generally. Cortical bone mass continued to be deficient in the
later period, indicating that stress was present throughout the Christian period, both early and
late. Unlike the long bone lengths that show a recovery during adolescence, there was a
continued decrease in cortical bone mass in older children, suggesting that growth in long
bone length continued at the expense of cortical bone maintenance (Hummert, 1983; and
compare with Garn et al., 1964).
Comparison of individuals interred in high-status brick vaults versus low-status grave pits
in a nineteenth-century setting in Birmingham, England revealed that the former had higher
cortical thickness for age than the latter (Mays, Brickley et al., 2009). These differences likely
reflect greater access to adequate nutrition in higher-status individuals than in lower-status
individuals in this early industrial setting.
Analysis of bone mass in living and past populations underscores an important point about
the plasticity of human skeletal tissue and its relationship with growth and development:
skeletal tissue adapts in a highly dynamic way to physiological requirements of the living
body. These requirements are informed by the environment, including biological, social, and
cultural influences over the life course of the individual, from conception through infancy,
childhood, adolescence, and adulthood (and see Agarwal & Beauchesne, 2011). This life course
perspective is motivated by the understanding that the skeleton reflects the lived experience of
the individual at all stages of life. This contextualized approach provides a powerful tool for
understanding variation in bone mass and the implication for risk of fracture and debilitation
(Glencross & Agarwal, 2011; and see Chapter 4).

2.3.2 Terminal adult height (stature)


As with bone mass, there is a substantial body of evidence drawn from the study of living
populations showing a strong relationship between growth suppression due to poor environ-
mental conditions during childhood and attainment of adult body size. In this regard, growth-
retarded children should develop into short-statured adults. The study of living populations
provides some support for this conclusion. Comparison of growth of undernourished Thai
children with American (US) children reveals that despite a longer period of growth in the
former (by about one year), the reduction in growth over their lifetimes resulted in shortened

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.3 Growth and development: skeletal 15

terminal height (Bailey et al., 1984; and see Bogin & MacVean, 1983; Frisancho, Newman
et al., 1970; Satyanarayana et al., 1980).
The close ties between environmental stress – especially due to poor nutrition – and adult
height are abundantly documented in research developing out of a growing interest in
anthropometric history (Floud et al., 1990; Heyberger, 2007; Komlos, 1994, 2009; Steckel,
1995, 2005; Steckel & Floud, 1997; Tatarek, 2006; Vercellotti et al., 2011; and many others).
Originally inspired by controversy over the health and well-being of enslaved African-
Americans (Steckel, 1979), current research has broadened greatly to include a range of other
populations in North America, Europe, and Asia (Fogel et al., 1983; Komlos, 1994; Steckel,
1995). Evidence from a wide range of recent historical populations indicates that stature
variability can be explained in large part by environmental factors (Steckel, 1995). This
evidence shows that terminal height is a product of nutritional adequacy and, to a lesser
extent, disease history. Individuals with adequate nutrition tend to reach their genetic growth
potential; those with poor nutrition do not.
Genetic factors are also important in determining terminal height. For example, well-off
Japanese reach only the fifteenth height percentile of well-off British (Tanner et al., 1982), and
genomic studies reveal associations with height (Becker et al., 2011). Climate may be a
mediating factor in determining height, but stature shows little correlation with latitude in
the comparison of a wide range of human populations (Ruff, 1994a). Of much more import-
ance to the issue of climate is body breadth, which plays a crucial role in determination of the
amount of body surface area to body mass in hot and cold climates (Ruff, 1994a).
Like childhood growth, there is a temporal trend of stature increase with economic and
nutritional improvement (Boldsen, 1995; Floud, 1994; Greulich, 1976; Komlos & Kriwy, 2002;
Malina et al., 2005; Őzer et al., 2011; Yagi et al., 1989; and many others) and decline during
times of hardship and deprivation (Bogin & Keep, 1999; Fogel et al., 1983; Kemkes-
Grottenthaler, 2005; Kimura, 1984; Leonard et al., 2002; Price et al., 1987; Steegmann,
1985). These analyses show that although growth and height have a genetic component,
environmental factors play a profound role in these outcomes.
Terminal height data for historical populations are drawn from various archival sources,
including military records (Bielicki & Welon, 1982; Komlos, 1989; Mokyr & Ó Gráda, 1994;
Sandberg & Steckel, 1987; Steegmann, 1985, 1986; Steegmann & Haseley, 1988), military
preparatory schools (Komlos, 1987), prison inmates (Riggs, 1994; Tatarek, 2006), enslaved
African Americans (Steckel, 1979, 1986, 1987), and voter registrations (Wu, 1994). Analysis
of these data sets by economic historians reveals temporal trends in stature that can be
linked with changing economic conditions relating to nutritional adequacy in particular and
health status in general. Terminal stature in Euroamerican populations shows significant
variability in relation to time, geography, and socioeconomic status. Over the last several
centuries, marked improvements in health and nutrition have been documented. Popular
convention holds that adult stature has increased during and after the Colonial period in
North America. Steckel (1994) analyzed stature data for American-born Euroamerican male
soldiers for the period of 1710 to 1950. Contrary to convention, twentieth-century Euro-
american males are not appreciably taller than their predecessors living in the eighteenth
century (Steckel, 1995).
Skeletons from archaeological contexts offer an important complementary data set for
stature analyses based on archival sources. Comparison of stature estimates derived from

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
16 Stress and deprivation during growth and development and adulthood

measurements of long bones shows little change from the pre-modern (1675–1879) to the
modern (1950–1975) period in the United States (Angel, 1976; Larsen, Craig et al., 1995)
(Table 2.1). Analysis of military archival records shows a slight rise in stature (1710–1830) in
European-descent, American-born males, followed by a marked decline that continues for
most of the nineteenth century (Costa & Steckel, 1997). This decline appears to coincide with
the movement of the population from a rural to urban setting, and increasingly poorer
sanitation. Simply, the height reductions reflected the health declines of the newly urbanized
people in the United States. Following the passage of sanitation laws in cities in the late
nineteenth century, improved nutrition, and healthier environments overall, there is a steady
rise in stature. The skeletal record for stature in the comparison of medieval and early modern
Europeans reveals a reduction in height, reflecting reduced quality of life, exposure to
pathogens, and decline in nutrition in the shift from rural to urban settings (Kemkes-
Grottenthaler, 2005).
In the New World, the transition to agriculture involved the adoption of maize as a key
component of subsistence. There are several negative aspects of maize that potentially could
lead to growth disruption and reduced height in native populations in the Americas. Although
maize meets energy requirements, it is deficient in the essential amino acids lysine, isoleucine,
and tryptophan (FAO, 1970; Whitney & Rolfes, 2011). Because maize has these amino acid
deficiencies, it provides a poor protein source. Niacin (vitamin B3) in maize is chemically
bound, which reduces the bioavailability of this nutrient to the consumer. In maize-based
diets, iron absorption is very low (Ashworth et al., 1973), methionine and phenylalanine are
minimally represented, and the leucine–isoleucine ratio is inadequate. The nutritive value of
maize is altered by the preparation techniques used to transform it into food. Many native
New World societies enhance the nutritional content of maize via alkali-processing (Katz
et al., 1974; Stahl, 1989). The addition of alkali promotes the availability of niacin during
digestion (FAO, 1953). Some evidence suggests that these treatment protocols actually pro-
mote dystrophic effects (Huss-Ashmore et al., 1982). Additionally, removal of the pericarp
(bran) in the grinding process decreases the nutritive value of maize; important minerals and
some fiber are removed if the pericarp is winnowed from the maize. If the aleurone, the
protein- and niacin-rich layer, and bran are removed simultaneously, important nutrients are
also lost (FAO, 1953; Rylander, 1994). Additionally, thiamine content is affected by the
manner in which the maize is processed.
The study of temporal series of archaeological remains, especially in the comparison of New
World foragers with later farming populations, reveals trends that are consistent with declin-
ing nutritional quality in both maize consumers and populations dependent on other plant
domesticates. Comparisons of prehistoric Georgia coastal foragers (pre-AD 1150) with later
maize farmers (AD 1150–1550) indicate reductions in stature of about 3% for adult females
and 1% for adult males (Larsen, 1982; Larsen, Crosby et al., 2002). Similar reductions in other
New World settings are documented in the American Midwest (Perzigian et al., 1984; Sciulli &
Oberly, 2002; but see Cook, 1984, 2007), Mesoamerica (Haviland, 1967; Márquez Morfín & del
Ángel, 1997; Nickens, 1976; Saul, 1972; Stewart, 1949, 1953a; but see Danforth, 1994;
Márquez Morfín et al., 2002; Wright & White, 1996), and Peru (Pechenkina, Vrandenburg
et al., 2007). Comparisons of agricultural populations with other settings indicate relatively
short statures in Mesoamerica (Storey, 1992a; Danforth, 1999), Ecuador (Ubelaker, 1994), and
Peru (Pechenkina, Vrandenburg et al., 2007), which are linked with chronic malnutrition.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.3 Growth and development: skeletal 17

Table 2.1 Euroamerican statures


Sample Description Dates N Stature1 Reference

Males
Patuxent Maryland rural 1658–1690 6 170 King & Ubelaker, 1996
Colonial-Civil War Various 1675–1879 21 173 Angel, 1976
Belleview Georgia rural 1738–1756 3 170 Rathbun & Scurry, 1991
Old Quebec Colonial military 1746–1747 30 173 Cybulski, 1988; Saunders, 1991
Walton Connecticut rural 1750–1830 5 176 Bellantoni et al., 1997
Fort William British military 1755–1757 14 177 Steegmann, 1986
Henry
Fort Laurens US military 1779 13 172 Sledzik & Sandberg, 2002
Bradford’s Colonial military 1812–1814 180 174 Saunders, 1991
Company
Snake Hill US military 1814 23 174 Sledzik & Sandberg, 2002
St. Thomas’ Ontario village 1820–1860 127 172 Saunders et al., 2002; Sledzik &
Sandberg, 2002
Prospect Hill Ontario 1824–1879 17 173 Pfeiffer et al., 1989
immigrants
Harvie Ontario rural 1825–1894 5 171 Saunders & Lazenby, 1991
Highland Park New York 1826–1863 94 172 Higgins et al., 2002
poorhouse
Cross Illinois rural 1829–1849 5 175 Larsen, Craig et al.,1995
Uxbridge Massachusetts 1831–1872 9 170 Sledzik & Sandberg, 2002;
poorhouse Wesolowsky, 1989;
Mt. Gilead Georgia rural 1832–1849 5 172 Wood et al., 1986
Voegtly Pennsylvania 1833–1861 32 170 Ubelaker & Jones, 2003
town
Glorieta Pass US military 1862 24 173 Sledzik & Sandberg, 2002
Little Big Horn US military 1876 8 176 Sledzik & Sandberg, 2002
West Point cadets2 US military 1840s–1870s 334 172 Komlos, 1987
Modern US2 General 2003–2006 2331 178 McDowell et al., 2008
population
Females
Patuxent Maryland rural 1658–1690 5 161 King & Ubelaker, 1996
Colonial-Civil War Various 1675–1879 7 160 Angel, 1976
Walton Connecticut rural 1750–1830 8 166 Bellantoni et al., 1997
Harvie Ontario rural 1825–1894 4 161 Saunders & Lazenby, 1991
Highland Park New York 1826–1863 64 161 Higgins et al., 2002
poorhouse
Cross Illinois rural 1829–1849 6 163 Larsen, Craig et al., 1995
Mt. Gilead Georgia rural 1832–1849 3 162 Wood et al., 1986
Voegtly Pennsylvania 1833–1861 14 160 Ubelaker & Jones, 2003
town
Modern US2 General 2003–2006 2477 163 McDowell et al., 2008
population
1
Mean values in cm.
2
Documentary/living statures.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
18 Stress and deprivation during growth and development and adulthood

1.5
Males

1 Females

0.5
Z-scores, mean+/-SE

-0.5

-1

-1.5

-2
Beiliu Jiangzhai Shijia Kangjia Xicun
Yangshao Longshan Dynastic

Figure 2.4 Variations in long bone lengths between Yangshao and Longshan cultures, as
expressed in average Z-scores. (From Pechenkina et al., 2002; reproduced with permission
of authors and John Wiley & Sons, Inc.)

In Mesoamerica, prehistoric heights are greater than in modern Maya (Danforth, 1999;
Márquez Morfín & del Ángel, 1997), which appears to be related to increased reliance on
maize and associated nutritional decline, especially in the last 500 years.
Other archaeological settings show reduction in stature in the shift to agricultural econ-
omies or agricultural intensification. Analysis of height data shows that foragers or incipient
farmers in settings in East Asia were taller and more robust than their farming descendants
(Bulbeck & Lauer, 2006; Kennedy, 1984; Temple, 2008; but see Clark et al., 2014). With regard
to millet farmers in the Wei and Yellow River area of China, earlier less intensive farming
populations are considerably taller than later more intensive farmers (Pechenkina et al., 2002,
2007, 2013) (Figure 2.4). Similarly, comparisons of skeletal series from the Upper Paleolithic
through the Neolithic in western Europe indicate a general reduction in average stature
(Bennike & Alexandersen, 2007; Meiklejohn et al., 1984; Meiklejohn & Babb, 2011; Roberts &
Cox, 2003, 2007; but see Jacobs, 1993). In some settings, reduction in stature coincided with
resource intensification, with either agriculture (Pechenkina et al., 2002, 2013; Van Gerven
et al., 1995) or foraging (Ginter, 2011). However, analysis of the record for Europe reveals
no change in the trajectory of stature reduction with the foraging-to-farming transition
(Meiklejohn & Babb, 2011). Nevertheless, the overall record suggests a link involving diet,
resource acquisition, physiological stress, and terminal height.
Much of the research on body size in children and adults in archaeological settings is
oriented toward tracking the consequences of adaptive transformations, primarily from
foraging to farming; relatively little is known about other dietary transitions. The

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.3 Growth and development: skeletal 19

consequences of change in dietary focus not involving agriculture are manifested in temporal
comparisons of native populations from the Santa Barbara Channel region of southern
California (Lambert, 1993; Walker, 2006; Walker & Thornton, 2002). In this region, popula-
tions shifted their dietary emphasis from terrestrial resources – especially plant foods – to
marine resources after 500 BC (Glassow, 1996). Over the period of 5500 BC to AD 1250, stature
reduced by about 7 cm (Walker & Thornton, 2002; and see Lambert, 1993, and compare with
Temple, 2008). Height reduction was fostered by decline in health due to the combined effects
of declining nutrition and elevated infectious disease. Protein, mostly derived from fish, was
abundant, but other important nutrients may have been lacking in the diets of later prehistoric
populations. During the latest prehistoric period (post-AD 1250), there was a rebound in
stature, possibly due to improved living conditions.
Similar trends in stature reduction have been documented in the Central Valley of interior
California (Ivanhoe, 1995; Ivanhoe and Chu, 1996). Comparisons of populations spanning the
period of 3000 BC to the mid-nineteenth century reveal statistically significant reductions in
stature for both females and males (2.2% and 3.1%, respectively). Archaeological evidence
indicates an increased reliance on acorns. Thus, stature reductions more likely reflect nutri-
tional stress owing to a focus on carbohydrates and a narrowing of the dietary spectrum in
later prehistory in this setting.
While the foraging-to-farming transition shows a general pattern of stature reduction
(Mummert et al., 2011), stature reductions identified in archaeological contexts are not
universal. A number of regions show either no change, an increase, or a high degree of
regional variability in stature (Danforth, 1994; Douglas & Pietrusesky, 2007; Roberts & Cox,
2007; Temple, 2011). In the lower Illinois River valley, there is no clear trend of stature change
in the comparison of early prehistoric through late prehistoric periods (Cook, 1984). This is
especially significant because it indicates that reduced juvenile height in this setting did not
result in reduced adult stature in later prehistoric agricultural groups. Likewise, temporal
comparisons of stature in a diversity of archaeological populations – from Ontario, northern
Great Plains, Peru, and Chile – show no change in stature with the shift in adaptive strategies
involving agriculture (Allison, 1984; various in Cohen & Crane-Kramer, 2007; Cole, 1994;
Katzenberg, 1992; Steckel & Rose, 2002). From c. 8250 yBP to the colonial period in Ecuador,
there is no evidence of stature decline despite increases in physiological stress (Ubelaker,
1994). All groups in Ecuador are relatively short-statured, thus stress (including poor nutri-
tion) may have been severe throughout the entire sequence (Ubelaker, 1994).
The influence of nutritional deprivation on human growth and terminal height and/or long
bone length is revealed in the study of components of past groups that may have been
differentially buffered against stress. Comparison of elite and non-elite adults from Old World
settings shows evidence of greater height in elites than in non-elites (Angel, 1975, 1984;
Becker, 1993; Buzon, 2006; Vercellotti et al., 2011). In a Maitas-Chiribaya (c. 2000 yBP)
population from northern Chile, shaman males are taller than other, non-elite males, which
may indicate better health and resources in the former (Allison, 1984). High-status adult males
in some Mesoamerican populations appear to be taller than low-status individuals or the
general population (Haviland, 1967; Helmuth & Pendergast, 1986–1987; Storey, 1998; but see
Wilkinson & Norelli, 1981). Likewise, elite males are taller than non-elite adults in several
contexts in prehistoric Southeast, Midwest, and Southwest United States (Buikstra, 1976a;
Cook, 1984; Hatch and Willey, 1974; Malville, 2008; Powell, 1988). These apparent status

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
20 Stress and deprivation during growth and development and adulthood

differences in attained height suggest that elites may have had nutritional advantages such as
having greater access to animal protein, resulting in achieving greater height than non-elite
individuals (Malville, 2008). Many of the elite distinctions are in adult males, suggesting that
the burden of stress may be primarily on adult males in ranked societies, at least as it is
exhibited in attained height.
This discussion points to the critical importance of understanding body size in relation to
resource availability and social context. Analysis of elite and non-elite individuals from San
Michele’s church in Trino Vercellese, Italy (c. AD 750–1300) shows considerable biological
variation in stature – high-status adults are taller than low-status adults – that is clearly
linked to relative social position in this Medieval setting (Porro et al., 1999). Additional
analysis of stature, body mass, and body proportions in this series reveals significant differ-
ences in high-status adults (Vercellotti et al., 2011). In particular, the comparison of different
body segments revealed significant differences between low- and high-status males. In
particular, males’ distal lower limb and body mass are significantly different in high- versus
low-status individuals. High-status men had longer distal lower limbs and were taller than
low-status men. However, the taller, high-status men had the shortest relative size of the
lower limb, indicating that height is not necessarily influenced by limb length. Rather, other
factors such as body trunk size may have a more important influence on height, at least for
members of a population having the most positive living conditions. The association between
stature and limb length in males is not shared by females in this setting. This may be due to
greater susceptibility to growth disruption in males than in females. The argument of greater
buffering in females than in males is also supported by the presence of greater prevalence of
enamel hypoplasias in men, especially low-status men, than in women in this setting.

2.3.3 Cranial base height


Biological anthropologists note specific patterns of variability in skull base height (auriculare–
basion or porion–basion distances) in selected samples, which Angel (1982) suggests is linked
to nutritional adequacy during the years of growth and development. Poorly nourished
individuals should have flatter cranial bases (called “platybasia”) than well-nourished indi-
viduals due to relatively greater deformation of supporting bone in response to the weight of
the head and brain: the “weakening of the bone from nutritional deficiencies decreases its
ability to resist gravitational pull, therefore inhibiting upward growth of the skull. . .Thus the
amount of compression in this area should give an indication of nutritional status” (Angel,
1982:298).
Angel tested his hypothesis by comparing skull base heights from skeletal series represent-
ing nutritionally disadvantaged and advantaged populations. These comparisons revealed
that the advantaged group has much higher cranial bases than the disadvantaged group,
which Angel concludes “fits a nutritionally caused mechanical weakening of bone supporting
a heavy head” (1982:302). The study of archaeological remains from the eastern Mediterra-
nean Basin indicates variation in cranial base height that Angel (1984; Angel & Olney, 1981)
attributed to nutritional quality: crania from populations experiencing nutritional deprivation
are platybasic, whereas crania from populations or segments of populations (e.g., Middle
Bronze Age “royalty”) with nutritionally adequate diets are not.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.3 Growth and development: skeletal 21

The link between cranial base height and nutritional quality may be more apparent than
real, however. Cranial base cartilages, like epiphyseal cartilages of limb bones, are primary
cartilages. Primary cartilages have intrinsic growth capabilities that are characteristically
resistant to compressive loading. This suggests that a model invoking compression as a causal
factor in determining cranial base form is incorrect. Therefore, the phenomenon of cranial
base flattening, while interesting, is largely unexplained.

2.3.4 Pelvic morphology


Severe vitamin D deficiency (rickets in children and osteomalacia in adults) caused by
prolonged lack of exposure to sunlight or inadequate intake of foods with vitamin D (e.g.,
eggs and oily fish) weakens maturing bone because the rapidly forming protein matrix does
not mineralize sufficiently. This results in pelvic deformation due to the forces created by
body weight and gravity (Angel, 1975, 1978a, 1982, 1984; Angel & Olney, 1981; Brickley &
Ives, 2008; Brickley et al., 2005, 2007; Greulich & Thoms, 1938; Nicholson, 1945; Thoms,
1947, 1956; Thoms et al., 1939; Walker et al., 1976). Pelvic inlet deformation is character-
ized by a reduction in the anterioposterior diameter relative to the mediolateral diameter
(called “platypellism”). Flattening of the pelvis is well documented in clinical populations
(Brickley & Ives, 2008; Greulich & Thoms, 1938; Nicholson, 1945; Thoms, 1947; Vieth,
2003) and in modern anatomical samples that compare lower- and middle-class groups
from the United States (Angel, 1982). For example, British women who were young children
during the war years of 1914–1918 have flattened pelvic inlets (Nicholson, 1945). Presum-
ably, these women had relatively poor nutrition during these years. Consistent with the
relationship between growth and nutritional status, women with flattened pelves also tend
to be short-statured.
Comparisons of pelvic inlet form between earlier and later (or modern reference) popula-
tions suggest improvements in nutritional health in several settings, including the eastern
Mediterranean (Angel, 1984; Angel & Olney, 1981), North America (Angel, 1976), and
Sudanese Nubia (Sibley et al., 1992). Preliminary evidence shows differences in pelvic shape
by social status group. Low-status adult females from the Middle Woodland (Klunk and
Gibson Mound groups) period in the lower Illinois River valley have flatter pelvic inlets than
high-status adult females (Brinker, 1985). These differences appear to reflect better nutrition
in the high-status women than in low-status women.
Other aspects of pelvic morphology may also be linked to negative environmental factors.
Sciatic notch widths are appreciably larger in nutritionally stressed eighteenth- and
nineteenth-century British from St. Bride’s Church, London, than in better fed twentieth-
century Americans (Walker, 2005). Rickets was a severe health problem in industrial England,
and is well documented in the St. Bride’s Church population. Archival documents indicate that
the births of St. Bride’s individuals with wide sciatic notches occurred during cold months of
the year, the period when rickets was especially prevalent. Analysis of a contemporary
population from Spitalfields, London, reveals that individuals with rickets had also been
exposed to extremely cold temperatures during the first year of life (Molleson & Cox, 1993).
Thus, the relatively wide sciatic notches in the St. Bride’s Church population appear to be
remnants of early childhood stress (Walker, 2005).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
22 Stress and deprivation during growth and development and adulthood

2.3.5 Long bone diaphyseal form


Pronounced bowing of the lower limb long bones is another skeletal deformation in rickets
and osteomalacia. As with the pelvis, most bowing deformities occur during the first several
years of life when the skeleton is undergoing rapid growth, especially between ages six
months and three years (Brickley & Ives, 2008; Stuart-Macadam, 1989). Thus, most femoral
diaphyseal deformations associated with severe vitamin D deficiency seen in adults likely
occurred during childhood. However, deformation can occur in adulthood, such as with
bending of the proximal femoral diaphysis and pelvic deformity (Brickley et al., 2007).
Vitamin D deficiency became highly prevalent during the Industrial Revolution, especially
in large, densely populated towns and cities in Europe. Culturally influenced avoidance of
sunlight (e.g., excess clothing, infant swaddling) may involve decreased vitamin D synthesis,
such as in Asia and North Africa (Fallon, 1988; Kuhnke, 1993). Increased availability of
vitamin D-enriched foods and reduced air pollution resulted in a virtual disappearance of the
disease in industrialized nations during the twentieth century.
Skeletal evidence of rickets is uncommon prior to the Medieval period in Europe. In
Medieval and later skeletal samples from Europe, a number of long bone deformities –
especially severe bowing of upper and lower limb bones – have a rachitic origin (Gejvall,
1960; Giuffra, Vitiello et al., 2013; Mays et al., 2006, Mays, Ives et al., 2009; Møller-
Christensen, 1958; Molleson & Cox, 1993) (Figure 2.5). Extreme bowing of lower limb bones
of an eight-year-old recovered from an early nineteenth-century African American cemetery
in Philadelphia probably resulted from rickets (Angel et al., 1987). A significant prevalence of
adult males and females – 35% and 20%, respectively – have bowing resulting from childhood
growth disturbance in the same population. Similar patterns of long bone bowing are
documented in a nineteenth-century African American series in North Carolina (Lambert,
2006) and in Old World contexts from the Iron Age site of Mahujhari, India (Kennedy, 1984),
and in Mesolithic, Bronze Age, and Medieval Europe (Giuffra, Vitiello et al., 2013; Meiklejohn &
Zvelebil, 1991). The study of long bones from a large cemetery in nineteenth-century
Birmingham, England, revealed a high prevalence of skeletal deformities due to rickets. The
condition contributed to reduced stature, especially for individuals over two years of age, due
to the combination of femoral bowing and endochondral bone deficiency (Mays, Brickley
et al., 2009). This investigation links lesion prevalence (rickets) and growth analysis.
Rickets is often associated with vitamin D deficiency in social contexts involving less
advantaged members of communities. The study of the remains of a very wealthy elite Medici
family of sixteenth-century Italy reveals, however, that children (newborn to five years old)
interred in the Basilica of San Lorenzo Medici burial crypt in Pisa, Italy, have deformed limb
bones, both upper and lower. These deformations in both upper and lower limbs reflect
modifications occurring when these children were first crawling and subsequently walking
(Giuffra, Vitiello et al., 2013). In addition to dietary deficiencies, historical sources indicate the
emphasis on swaddling – wrapping infants in heavy layers of cloth – and sequestration in
homes well away from sunlight. Thus, even in the most elite social classes and in a setting of
the world with considerable sunlight, elite children of highly advantaged members of society
were at considerable risk for developing rickets.
Prehistoric evidence of rickets in the Americas is exceedingly rare, but a pattern of femoral
and tibial shaft deformation in at least one setting in the American Southwest has been

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.3 Growth and development: skeletal 23

Figure 2.5 Rickets in left juvenile (four to five years old) ulna (left) and radius (right);
Medieval period, Gruczno, Poland. (From Kozłowski, 2012; photograph by Tomasz
Kozłowski.) (A black and white version of this figure will appear in some formats. For the
color version, please refer to the plate section.)

identified. This appears to be associated with a segment of a late prehistoric society that had
experienced relatively elevated levels of physiological stress generally (Palkovich, 2008).
Flattening of femoral and tibial diaphyses has been documented in numerous archaeo-
logical skeletal samples worldwide (see Chapter 6). The primary indices measuring the degree
of flatness of femora and tibiae include the meric index (anterioposterior flattening of
the proximal femoral diaphysis), pilasteric index (mediolateral flattening of the femoral

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
24 Stress and deprivation during growth and development and adulthood

diaphysis), and cnemic index (mediolateral flattening of the tibial diaphysis at the nutrient
foramen). Some attribute diaphyseal flattening to nutritional stress (Adams, 1969; Angel,
1984; Buxton, 1938). Buxton (1938) asserted that less bone is required in the construction of
a diaphysis if it is flattened rather than round. He viewed the temporal trend of rounder
diaphyses as representing an increase in amount of bone, inferring a decline in nutritional
deficiency in recent “civilized” populations. Structural analysis of long bone diaphyses
reveals that flattening is related not to the amount of bone present, rather to the manner
in which it is distributed when viewed in cross-section (Ruff, 2008). Mechanical loading, not
nutritional stress, is the primary determinant of flatness of long bone diaphyses (see
Chapter 6). Nutritional deprivation or other physiological stressors certainly have an influ-
ence on amount of bone, but the relationship between nutritional status and diaphyseal
shape is unsubstantiated.

2.3.6 Vertebral neural canal size


The effects of catch-up growth on stature and long bone lengths are problematical for
documenting the stress history of an individual during their growth years. An individual
may be stressed early in life, but amelioration of negative conditions (e.g., improvement in
nutritional status) during later juvenile years may result in obliteration of evidence of
growth disruptions that had occurred earlier in life. In the Dickson Mounds, Illinois series,
for example, although juvenile growth became stunted in the transition to intensive
farming for the period of AD 950 to 1300, no appreciable reductions occurred in adult
height (Lallo, 1973). Thus, adult heights in this population are uninformative about
juvenile stress.
The temporal similarity of stature in the Dickson Mounds series may be simply due to
growth recovery. Vertebral growth provides a means of addressing the problem of growth
stress identification not possible with attained height. At the time of birth, vertebral neural
canal size is approximately 65% complete; full size is reached by about four years of age
(Clark, 1988; Clark et al., 1986). Vertebral body height continues to grow into early adulthood,
well after the third decade of life. Thus, early and late stress in the life history of the individual
is represented in the respective size of the vertebral neural canal and vertebral body height in
adult skeletons. If there is a reduction in canal size but not in vertebral height, then catch-up
growth likely occurred following early stress (prior to four years of age). If both neural canal
size and vertebral body height are small, then stress was likely present throughout most of the
years of growth and development, certainly after four years of age and possibly into adult-
hood (Clark, 1988; Clark et al., 1986).
Analysis of thoracic and lumbar vertebrae from the Dickson Mounds site reveals that
growth of neural canal size was completed prematurely, but growth in vertebral body
height continued through the juvenile years into adulthood (Clark et al., 1986). This
growth pattern suggests that stress amelioration in later childhood accounts for the
similarity in adult long bone lengths and stature in the earlier and later populations from
the Dickson Mounds site.
Young adults (15 – 25-year age group) in the Dickson Mounds series have a significantly
smaller vertebral neural canal size than older adults (25þ years) (Clark et al., 1986).
Similarly, in the Medieval-period Fishergate House series, United Kingdom, individuals aged

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.4 Growth and development: dental 25

17 – 25 years have a significantly smaller vertebral neural canal size than older adults (Watts,
2011). This size variation suggests a link between skeletal development and conditions that
predisposed them to earlier death. Simply, individuals stressed during the juvenile years of
growth and development are likely to die earlier than individuals who are either not stressed
or less stressed.

2.4 Growth and development: dental


2.4.1 Dental development rate
Dental development is comprised of two components: formation of crowns and roots and
eruption of teeth. Dental development overall is less sensitive to environmental constraints
than is skeletal development (Cardoso, 2007; Demirjian, 1986; Smith, 1991). The relatively
greater resistance of dental tissues to environmental insults has been demonstrated by the
observation that various stressors influencing stature and bone age have a relatively small
effect on dental development (reviewed in Cardoso, 2007; Smith, 1991). The high degree of
genetic control over dental development serves to minimize the effects of poor environmental
circumstances (Cardoso, 2007; Garn et al., 1965; Holman & Yamaguchi, 2005; Moorrees &
Kent, 1981; Smith, 1991).
Tooth formation rates are relatively free of environmental influence (e.g., nutrition), which
is indicated by low correlations between tooth formation and skeletal age, stature, relative
body weight, fatness, and by the lack of any kind of secular trend (Smith, 1991). Eruption
rates and timing, however, are more responsive to environmental factors, such as caries
experience, tooth loss, and especially poor nutritional status (Alvarez, 1995; Alvarez & Navia,
1989; Alvarez et al., 1988, 1990; Gaur & Kumar, 2012; Holman & Yamaguchi, 2005; Oziegbe
et al., 2014; Ronnerman, 1977). For example, eruption and exfoliation of deciduous teeth are
delayed in nutritionally deprived children in comparison with well-nourished children from
Cantogrande, Peru (Alvarez, 1995; Alvarez et al., 1988; and see Barrett & Brown, 1966;
Cardoso, 2007). Additionally, unlike formation, eruption timing shows some correlation with
body size (Garn et al., 1960; McGregor et al., 1968).
It is not possible to identify delays in dental development timing in archaeological series
based on teeth alone, because age-at-death must be determined by comparing the archaeo-
logical dentitions with some standard based on individuals of known age (Moorrees et al.,
1963). However, relative differences between dental and skeletal development may provide
some insight into growth stress. Comparison of skeletal age and dental age in Medieval-period
skeletons from Sudanese Nubia reveals that most individuals (70.5%) have skeletal ages that
are younger than their dental ages (Moore et al., 1986). These relative differences indicate that
skeletal growth may have been retarded. Dietary reconstruction suggests that growth retard-
ation may have been due to nutritional deprivation, a finding that is consistent with other
skeletal indicators of stress (e.g., porotic hyperostosis: Moore et al., 1986). Analysis of skeletal
and dental development where age is known reveals that dental development is less subject to
environmental perturbations than is skeletal development. For example, Cardoso’s (2007)
analysis shows that individuals with lower socioeconomic status in an urban setting in
Portugal have more delayed skeletal development than dental development, by as much as
half a year.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
26 Stress and deprivation during growth and development and adulthood

2.4.2 Tooth size


Like bone size, tooth size involves a complex interplay between environment and heredity.
Unlike skeletal elements, tooth crowns do not remodel once fully formed. Thus, teeth provide
a record of size well in advance of the adult years. Tooth size appears to be highly heritable,
indicating that variation between and within human populations can be explained mostly by
genetic differences (see Chapter 7; and see Kieser, 1990). Twin studies reveal that as much as
80% to 90% of observed variation in tooth size is due to genetic influence for anterior teeth
and 60% for posterior teeth (Townsend et al., 1994). Other estimates of heritability vary widely
(Kieser, 1990), but most workers agree that environmental influences on tooth size are
significant, albeit small (Dempsey et al., 1996; Dempsey & Townsend, 2001; Garn et al.,
1965; Garn, Osborne et al., 1979; Hughes & Townsend, 2013; Potter et al., 1983; Townsend,
1980, 1992; Townsend & Brown, 1978). Thus, tooth size represents a measure of deviation
from genetic growth potential in response to some stressor or stressors (Apps et al., 2004;
Bailit et al., 1968, 1970; Evans, 1944; Garn, Osborne et al., 1979; Garn et al., 1980; Goose,
1967). Placental insufficiency, low birth weight, maternal health status, nutritional status, and
a variety of genetic and congenital defects (e.g., Down’s syndrome, cleft palate, prenatal rubella,
congenital syphilis) are linked with reduced tooth size (Apps et al., 2004; Cohen et al., 1979; Garn
& Burdi, 1971; Garn, Osborne et al., 1979; Goodman, 1989; Seow & Wan, 2000). Understanding
the influence of nutrition on tooth size is hampered by the paucity of data holding genetic factors
constant in situations of variable nutritional quality. These findings are consistent with
experimental research on laboratory animals showing tooth size reduction in response to
developmental disruptions and nutritional deprivations (Bennett et al., 1981; Holloway et al.,
1961; Paynter & Grainger, 1956; Riesenfeld, 1970; but see Murchison et al., 1988).
Prehistoric maize agriculturalists from coastal Georgia post-dating AD 1150 have smaller
teeth than did their foraging predecessors (Larsen, 1982, 1983). Tooth size reduced in both the
permanent and deciduous dentitions, which may reflect an increase in physiological stress due
to declines in dietary quality and health status generally. Tooth size reduction in the primary
dentition suggests a negative change in maternal health status and placental environment as
deciduous teeth form in utero. Given the relatively narrow temporal window of tooth size
reduction in this and other populations with the shift from foraging to farming or farming
intensification (Bulbeck & Lauer, 2006; Christensen, 1998; Coppa et al., 1995; Hinton et al.,
1980; Meiklejohn & Zvelebil, 1991; Pinhasi & Meiklejohn, 2011; y’Edynak, 1989), these
changes likely indicate an increase in stress that accompanied this transition. In contrast,
Lunt (1969) documented a temporal increase in permanent tooth size from Medieval times to
the present in Denmark, which is attributed to improved dietary conditions in later times (and
see Lavelle, 1968).
Dental size decrease or increase in Holocene populations cannot be explained fully by non-
evolutionary factors. In prehistoric Nubian populations, there is a relatively greater reduction
in posterior tooth size than in anterior tooth size, which Calcagno (1989) attributes to a
selective advantage for smaller posterior teeth in caries-prone agriculturalists. These findings
underscore the complexity of tooth size, requiring consideration of both extrinsic and intrin-
sic circumstances in specific settings.
The hypothesis that members of a population who suffer most from illness and physio-
logical stress are more likely to die at an earlier age than are other (healthier) members of a

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.4 Growth and development: dental 27

Table 2.2 Juvenile and adult permanent tooth size (buccolingual; mm) from Santa Catalina de
Guale, St. Catherines Island, Georgia (Adapted from Simpson et al., 1990; Table 5-1)
Juvenile Permanent

N Mean SD N Mean SD % Differencea


Tooth
Maxillary
I1 16 7.66 0.56 33 7.48 0.40 −2.4
I2 23 6.94 0.39 37 6.91 0.36 −0.4
C 28 8.59 0.66 55 8.64 0.47 0.6
PM1 34 10.12 0.59 70 10.09 0.49 −0.3
PM2 25 9.77 0.56 72 9.89 0.64 1.2
M1 38 11.93 0.68 77 12.14 0.51 1.7
M2 21 12.09 0.67 85 12.01 0.68 −0.7
Mandibular
I1 20 5.84 0.38 22 5.89 0.33 0.8
I2 27 6.23 0.40 47 6.34 0.38 1.7
C 32 7.51 0.57 77 7.85 0.53 4.3c
PM1 37 8.09 0.46 95 8.30 0.44 2.5b
PM2 33 8.42 0.52 95 8.63 0.47 2.4b
M1 45 11.11 0.49 72 11.24 0.52 1.2
M2 31 10.76 0.57 87 10.76 0.61 0.0
a
Computed by the formula: 100 − [100  (min. mean/max. mean)].
b
P0.05 (Student’s t-test).
c
P0.01 (Student’s t-test).

population has been tested by the comparison of permanent tooth size of juveniles and adults
in different settings in the American Southeast, namely in the late prehistoric Averbuch series
from the Middle Tennessee River valley (Guagliardo, 1982a) and colonial Spanish missions
(Simpson et al., 1990; Stojanowski et al., 2007). Both populations were sedentary maize
agriculturalists exhibiting skeletal evidence of high levels of physiological stress and poor
health. In these settings, juveniles have smaller permanent teeth than adults. In the Santa
Catalina series, nine of 14 tooth types examined are smaller in juveniles than in adults
(Table 2.2). The other tooth types show either no difference or slightly smaller size in adults.
All statistically significant differences are for smaller juvenile teeth. In these samples, juvenile
– adult size differences are more common in mandibular teeth than in maxillary teeth,
suggesting that the lower dentition may be more developmentally sensitive to stress than is
the upper dentition.
These studies indicate the failure of teeth to reach their growth potential in circumstances
involving increased stress. This conclusion lends support for Sagne’s (1976) hypothesis that
the Medieval-era Swedish dying young received suboptimal nutrition during the years of
dental development, resulting in smaller teeth (and see Lunt, 1969). These investigations
suggest that individuals with small teeth – those who were most stressed – had a reduced
lifespan, which is consistent with evidence based on vertebral neural canal size, height

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
28 Stress and deprivation during growth and development and adulthood

(Kemkes-Grottenthaler, 2005), and other dental indicators (e.g., enamel defects: Goodman &
Armelagos, 1988). As with neural arch canal size and height, it is unlikely that small teeth led
to reduced longevity. Rather, size reduction is symptomatic of environmental stress that
contributed to smaller teeth and early death.

2.4.3 Fluctuating and directional asymmetry


Beginning with the work of developmental geneticists in the 1940s, a consensus has
emerged that bilateral structures normally developing as mirror images of each other will
develop differently in response to environmental instability (Kieser, 1990). Van Valen
(1962) suggested that one type of asymmetry – which he called fluctuating asymmetry –
reflects that when confronted with stress, the body tissues are unable to develop bilat-
erally in their normal growth pathways. Thus, in settings involving elevated stress, teeth
and other bilateral structures fail to develop evenly on both sides. Support for this
hypothesis is provided by the study of laboratory animals exposed to induced stress
(e.g., hypothermia, blood loss, heat, cold, diabetes, audiogenic stress). Stressed animals
display an increase in fluctuating asymmetry in a variety of bilateral anatomical struc-
tures, including dental and skeletal tissues (Albert & Greene, 1999; Klingenberg, 2003;
Kohn & Bennett, 1986; Møller & Swaddle, 1997; Nass, 1982; Palmer & Strobeck, 2003;
Polak, 2003; Richtsmeier et al., 2005; Sciulli et al., 1979; Siegel & Mooney, 1987; Siegel
et al., 1977).
The study of odontometric fluctuating asymmetry in living, archaeological, and paleon-
tological samples presents mixed and sometimes contradictory results (Barrett et al., 2012;
Bassett, 1982; Black, 1980; Corruccini et al., 2005; Doyle & Johnston, 1977; Guatelli-
Steinberg et al., 2006; Harris & Nweeia, 1980; Kieser & Groeneveld, 1998; O’Connell,
1983; Suarez, 1974; Townsend & Brown, 1980; Townsend & Farmer, 1998). Left–right
tooth size differences are present in archaeological samples, including the Archaic-period
Indian Knoll (Kentucky), late prehistoric Campbell site (Missouri), and contact-era Larson
site (South Dakota) (Perzigian, 1977). The Indian Knoll dentitions are the most asymmetric,
which Perzigian (1977) attributes to poor diet in comparison with late prehistoric and
contact-era agriculturalists. His interpretation of decrease in dietary quality in comparing
prehistoric foragers and farmers runs counter to the conclusions drawn by many bioarch-
aeologists working in the Eastern Woodlands – namely, nutritional stress is more pro-
nounced in farmers than in foragers (Larsen, 1995). Thus, the pattern of decreasing
asymmetry in these groups remains unexplained. Temporal comparisons of a series of
populations from prehistoric Paloma, Peru reveals a trend of decrease in asymmetry over
time, and in this setting, substantive skeletal and dental evidence indicates improving
health over time (Benfer, 1984).
Using simulation sampling, Smith and coworkers (1982; and see Garn, Cole et al., 1979)
assert that the amount of asymmetry is highly sensitive to sample size. They argue that sample
sizes of several hundred individuals are required in order to detect meaningful differences
between populations in fluctuating asymmetry. Similarly, Greene (1984) found that the
confounding effect of measurement error can both obscure real differences and artificially
create others.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.4 Growth and development: dental 29

Kieser and coworkers (1986a, 1986b; and see Kieser, 1990) analyzed fluctuating dental
asymmetry as an indicator of environmental disruption in highly stressed Lengua Indians
presently inhabiting the Chaco region of Paraguay. Application of Euclidean map analysis,
a statistically powerful approach whereby each dentition is considered an independent
variable, produces a measure of asymmetry represented by dividing the sum of Euclidean
distances for tooth antimeres by a product of the mean individual tooth size and the
number of tooth pairs. Comparisons with well-nourished, disease-free Whites reveals much
greater asymmetry in the Lengua population. Younger, more acculturated Lengua with
better diets and greater access to Western healthcare show lower asymmetry values than
those found in more traditional Lengua, experiencing elevated stress. Using the same
methodology, Townsend and Farmer (1998) determined asymmetry scores in a sample of
South Australian children. Most children are healthy, and have correspondingly low
asymmetry scores. A few individuals with low birth weight have relatively high asymmetry
scores.
Similar patterns of fluctuating asymmetry as a biological indicator of developmental
instability have been documented in other hard tissues. Based on the earlier finding that
environmental stress was likely higher in the early Christian period than in the late Christian
period in Nubia (and see earlier), DeLeon (2007) tested the hypothesis that there is a decrease
in fluctuating asymmetry in the craniofacial skeleton. Her analysis of landmark coordinate
data revealed a pronounced decrease in level of fluctuating asymmetry in the later popula-
tion compared to the earlier population (Figure 2.6). Interestingly, the lateral landmarks
show more asymmetry than that seen in the medial landmarks, supporting the notion that
developmental instability is trait specific and not random. Similarly, epiphyseal union in the
earlier, more highly stressed population shows considerably more bilateral asymmetry than
that found in the later population (Albert & Greene, 1999), also consistent with other
bioarchaeological evidence for decline in stress in the temporal sequence in Nubia. The
presence of fluctuating asymmetry in dental and craniofacial tissues would suggest the
likelihood of poor health outcomes in adults. Indeed, analysis of nineteenth- and early
twentieth-century remains from Portugal, having known cause-of-death, reveal that adults
who died from degenerative diseases (e.g., chronic heart disease, diabetes) had higher rates
of fluctuating asymmetry than those with other causes of death (e.g., infectious disease)
(Weisensee, 2013).
Directional asymmetry is another pattern of bilateral variation that has been identified in
analysis of tooth size in human populations. This pattern is characterized by larger teeth on
one side of the dental arch than on the other. Directional asymmetry is infrequently reported
for human populations (Ben-David et al., 1992; Boklage, 1987; Harris, 1992; Lundström,
1948; Mizoguchi, 1986; Sharma et al., 1986; Townsend & Farmer, 1998). Harris (1992)
detected directional asymmetry in a large sample of permanent teeth of Euroamerican
adolescents, with a consistently greater left dominance in one dental arch.
Directional asymmetry is unexplained by current models, but may be an indicator of
developmental instability arising from stress (Harris, 1992). The strong environmental basis
of directional (and fluctuating) asymmetry is inferred by observation of low intraclass
correlations between monozygous and dizygous twins (Townsend, 1992). Additionally,
detection of spurious genetic variance indicates a virtual lack of evidence for a significant
genetic basis.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
30 Stress and deprivation during growth and development and adulthood

EARLY PERIOD LATE PERIOD

LAM

PTP
NAS
AST

GPF

LAM

PTP

NAS
AST

GPF

NAS

PTP

AST

LAM

Figure 2.6 Fluctuating asymmetry observed in Early and Late Period Kulubnarti, Sudan.
More fluctuating asymmetry is observed in the Early Period sample, indicating elevated
physiological stress. (From DeLeon, 2007; reproduced with permission of author and John
Wiley & Sons, Inc.)

2.5 Skeletal and dental pathological markers of deprivation


2.5.1 Anemia and abnormal cranial porosities
Anemia (“without blood”) is a pathological deficiency in red blood cells or hemoglobin, the
key protein that chemically binds with oxygen, transporting it to the body tissues. Thus, with
a deficiency in either red blood cells or hemoglobin, an individual has a reduced availability of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 31

oxygen. The outcome of such a deficiency and associated compromised access to oxygen is
reduced physical activity in general and work performance in particular (Crouter et al., 2012;
Haas, 2006).
There are three causes of anemia – blood loss, impaired production of red blood cells,
increased red blood cell production, or any combination of these conditions. Some of the
anemias are genetic, such as thalassemia, sickle cell anemia, nonspherocytic hemolytic
anemia (e.g., glucose-6-phosphate dehydrogenase deficiency [favism], pyruvate kinase defi-
ciency), spherocytosis, and rarely, hereditary elliptocytosis. In living and past populations, the
majority of the anemias are acquired, caused by either blood loss or nutritional deficiencies.
There are regions of the world, however, where genetic anemias are highly prevalent, reflect-
ing natural selection in areas where malaria is endemic. Angel (1966a) made the case that
pathological conditions he documented in the Mediterranean basin were largely genetic in
origin. DNA sequence analysis from an eight-year-old from the site of el Akhziv, Israel,
revealed that the child had homozygosity for ß-thalassemia (Filon et al.,1995), thus providing
some support for Angel’s hypothesis. However, it now seems unlikely that the record of
cranial porosities in the region is largely due to genetic causes (and see later).
In the normal individual, the rate of red blood cell production is balanced by the rate of red
blood cell destruction. The balance between production and destruction of red blood cells
requires a number of key micronutrients, including especially the essential amino acids, iron,
and vitamins A, B12, B6, and folic acid (Martini & Ober, 2001; Walker et al., 2009). Iron is an
essential element of hemoglobin, thus serving as a key component of oxygen transport. Iron
deficiency is the most common cause of anemia, owing to blood loss, iron-poor diets, and
gastrointestinal malabsorption of the element.
The absorption of iron is dependent upon its source, from either heme or non-heme foods.
Generally, heme sources of iron are efficiently absorbed, with meat being among the best
(Baynes & Bothwell, 1990). Iron in meat does not require processing in the stomach, and the
amino acids derived from digestion of meat help to enhance iron absorption (Wadsworth,
1992). Iron bioavailability of non-heme sources is highly variable, but plant sources are
generally poorly absorbed. Various substances found in plants inhibit iron absorption, such
as phytates in many nuts (e.g., almonds, walnuts), cereals (e.g., maize, rice, whole wheat
flour), and legumes (Baynes & Bothwell, 1990). Unlike protein in meat, plant proteins (e.g., in
soybeans, nuts, lupines) inhibit iron absorption. Tannates found in tea and coffee also reduce
iron absorption significantly (Hallberg, 1981).
A number of foods enhance iron bioavailability, especially when consumed in combination
with non-heme iron. For example, ascorbic acid promotes iron absorption (Baynes &
Bothwell, 1990; Hallberg, 1981; Wadsworth, 1992). Citric acid from various fruits and lactic
acid from fermented cereal bees are implicated in promoting iron absorption (Baynes &
Bothwell, 1990). Non-heme iron (e.g., in maize) is enhanced considerably by concurrent
consumption with meat and fish (Layrisse et al., 1968; Navas-Carretero et al., 2008).
If either absorption or consumption of iron or the other micronutrients is low, red blood cell
production is reduced, resulting in anemia. In response to anemia, the body first steps up its
production of red blood cells (Stockmann & Fandrey, 2006). If this first response fails in
alleviating the shortfall, the regions of the skeleton responsible for red blood cell production
are activated, resulting in an expansion of the cranial vault marrow (diploë). This elevated red
blood cell production occurs at the expense of the ectocranial surfaces, resulting in resorption

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
32 Stress and deprivation during growth and development and adulthood

(a)

Figure 2.7a Porotic hyperostosis on frontal and parietals; Santa Maria de los Yamassee, Amelia
Island, Florida. (Photograph by Mark C. Griffin.) (A black and white version of this figure
will appear in some formats. For the color version, please refer to the plate section.)

of the bone and a characteristic porosity called porotic hyperostosis, especially involving
the cranial vault bones (Figure 2.7).
Marrow expansion, therefore, is clearly caused by anemia as a response to deficiencies in
red blood cells and/or hemoglobin (Moseley, 1974; Ortner, 2003; Schultz, 2001). However,
only some forms of anemia result in the characteristic pathognomic porotic lesions of the
cranial vault. That is, only the anemias that can sustain high and elevated production of red
blood cells likely cause porotic hyperostosis. Walker and coworkers (2009) argue that iron
deficiency anemia per se results in decreased production of red blood cells, anemia certainly,
but not the kind of anemia that could produce cranial porosities. That is, iron-deficiency
anemia is a condition of reduced and inadequate blood production and is insufficient, thus
resulting in the kind of marrow hypertrophy associated with porotic hyperostosis and cribra
orbitalia. Specifically, two forms of anemia lead to elevated red blood cell production and
cranial marrow expansion resulting in porotic hyperostosis, namely megaloblastic and hemo-
lytic anemias. Megaloblastic anemia is commonly caused by chronic nutritional deficiencies
and malabsorption of vitamin B12 or folic acid, or both. In both forms of anemia, there is an
overproduction of red blood cells, resulting in expansion of the marrow (Walker et al., 2009).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 33

(b)

Figure 2.7b Histological section from Figure 2.7a. The linear orientation and general
morphology of diploic cavities are consistent with iron deficiency anemia. (Schultz et al.,
2001; reproduced with permission of authors and University Press of Florida.) (A black and
white version of this figure will appear in some formats. For the color version, please refer
to the plate section.)

Similar lesions found in the roof areas of the eye orbits, called cribra orbitalia, are also
frequently observed in archaeological remains (Figure 2.8). There is an extensive body of
bioarchaeological research documenting porotic hyperostosis, largely beginning with Angel’s
(1966a, 1967, 1971a, 1978b, 1984) systematic study of a large series of skeletal remains on a
regional and population basis. Based on his study of some 2200 archaeological crania from
the eastern Mediterranean region, principally Greece, Cyprus, and Turkey, Angel proposed
that porotic hyperostosis resulted from hereditary hemolytic anemias, especially thalassemia
or sickle cell anemia. In this setting where malaria is endemic, individuals who are heterozy-
gous for sickle cell anemia or thalassemia have a selective advantage over normal homozy-
gous individuals lacking the sickle-cell or thalassemia genes. Carriers show lower infection
rates by malarial parasites (genus Plasmodium), thus enjoying greater protection from
malaria.
Other regional studies dealing with large samples of skeletal remains showed that cranial
porosities in past populations are likely due to a variety of nongenetic factors. In Wadi Halfa,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
34 Stress and deprivation during growth and development and adulthood

Figure 2.8 Cribra orbitalia; Santa Catalina de Guale de Santa Maria, Amelia Island, Florida.
(From Larsen, 1994; photograph by Mark C. Griffin; reproduced with permission of John
Wiley & Sons, Inc.) (A black and white version of this figure will appear in some formats.
For the color version, please refer to the plate section.)

Nubia, in the Nile Valley, high prevalence of orbital lesions (21.4%) for the Meroitic (350 BC –
AD 350), X-group (AD 350 – 550), and Medieval Christian (AD 550 – 1400) periods have been
reported (Carlson et al., 1974). Reconstruction of the environmental context based on arch-
aeological, historical, and ethnographic evidence indicates that several factors likely contrib-
uted to anemia. Milled cereal grains (millet, wheat), the focus of diet in this setting, contain
very little iron and are high in phytate. Additionally, as with populations currently living in
the Nile Valley, hookworm disease and schistosomiasis were likely highly endemic. These
factors, combined with chronic diarrhea that is also highly prevalent in the region today, are
consistent with the presence of anemia.
Further to the south in the Nile Valley, a high prevalence of porotic lesions (45%) has been
reported at the Medieval-period Kulubnarti site (AD 550 – 1500) (Mittler & Van Gerven, 1994).
Early and late-period Kulubnarti juveniles have remarkably high prevalences (94% and 82%,
respectively: Van Gerven et al., 1995). Like the Nubian groups down river at Wadi Halfa, the
Kulubnarti population suffered the ill effects of anemia owing to nutritional deficiencies and
other negative influences of sedentism and unhealthy living conditions. Analysis of demo-
graphic profiles of individuals with and without lesions indicates that those with porotic
crania are associated with a shortened life expectancy, with differences greatest in the
subadult years (and see Blom et al., 2005). There is a decline in porotic lesion prevalence
from 51.8% to 39.0% from the early to late Christian periods (AD 550 – 750 to AD 750 –
1500). This apparent improvement in iron status coincides with improvements in health

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 35

generally that arose following increased political autonomy and improved living conditions
(Mittler & Van Gerven, 1994).
Circumstances involving association between anemia and increasing or elevated stress
have been documented in Medieval and seventeenth-century Tokyo (Hirata, 1990), prehistoric
Iran and Iraq (Rathbun, 1984), third-century BC Carthage (Fornaciari et al., 1981), Neolithic
Levant and Greece (Eshed et al., 2010; Papathanasiou, 2005; Papathanasiou et al., 2000),
third-century AD Moldavia (Miritoiu, 1992), and Romano-British, Medieval, and eighteenth–
nineteenth-century Britain (Grauer, 1993; King et al., 2005; Lewis, 2002; Molleson & Cox,
1993; Stuart-Macadam, 1991; Sullivan, 2005; Wells, 1982). All of these settings have good
contextual evidence for elevated environmental stress, but the particular circumstances for
anemia are regional specific. For example, causative factors for high prevalence of porotic
lesions in the Roman-period Poundbury Camp, and likely in British populations, included
parasitism, infectious disease, and perhaps lead poisoning (Stuart-Macadam, 1991). High
prevalences in eighteenth–nineteenth-century urban London appear to be linked with shifts
in weaning practices, poor living conditions, and low maternal health status (Lewis, 2002;
Molleson & Cox, 1993). Indeed, comparisons of temporal series in England reveal that
industrialization had the highest negative impact on children’s health, more so than events
in any other period (Lewis, 2002). On the other hand, improved environments result in the
decline of prevalence of porotic lesions. For example, decrease in prevalence in modern Japan
reflects decreased crowding, reduction in infectious diseases, and improved hygiene (Hirata,
1990; Temple, 2010).
The most abundant data on porotic lesions is available from the New World, especially
North America. In the American Southwest, cranial porosities are highly prevalent, especially
in late prehistoric and historic-era populations (Akins, 1986; El-Najjar & Robertson, 1976; El-
Najjar et al., 1975, 1976, 1982; Hooton, 1930; Kent, 1986; Martin et al., 2001; Palkovich,
1980, 1987; Stodder, 1994, 2006; Stodder & Martin, 1992; Stodder et al., 2002; Walker, 1985;
Walker et al., 2009; Zaino, 1967, 1968). Among mostly late prehistoric Puebloan samples
studied by El-Najjar and collaborators (El-Najjar et al., 1976) from Canyon de Chelly, Chaco
Canyon, Inscription House, Navajo Reservoir, and Gran Quivira, porotic lesions were docu-
mented in 34.3% of individuals. At Chaco Canyon alone, some 71.8% of individuals display
the characteristic lesions. Similarly, high prevalences have been reported from late prehis-
toric- and contact-period sites, including San Cristobal (90%), Hawikku (84%), Black Mesa
(88%), Mesa Verde (70%), Dolores (82%), Cases Grandes (46%), and La Plata Valley (40%)
(Stodder, 1994; Weaver, 1985). There are some Southwestern samples that have relatively low
prevalences (e.g., 16% for Navajo Reservoir children; see Martin et al., 1991). Martin and
coworkers (1991) note that comparisons of data collected by different researchers is problem-
atical because of the varying methods used in identification and recording of porotic lesions.
For example, some researchers may include slight pitting when analyzing their data sets,
whereas others may not. Unfortunately, this distinction is only rarely noted in bioarchaeo-
logical reports, regardless of geographic or cultural setting.
El-Najjar (1976) links the elevated levels of abnormal cranial porosities in the American
Southwest and other regions of the New World to the effects of over-reliance on maize in
conjunction with food processing techniques that may contribute to iron deficiency. Specif-
ically, he regards the presence of phytate – an iron inhibitor – as well as lime treatment to
decrease the nutritional value of maize.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
36 Stress and deprivation during growth and development and adulthood

Analysis of archaeological samples from other maize agriculturalists in the New World
provides mixed support for El-Najjar’s dietary hypothesis. Relatively high prevalences of
porotic lesions (>15% – 20%) are present in agriculturalists in the American Midwest (Cook,
1984; Garner, 1991; Goodman, Lallo et al., 1984; Lallo et al., 1977; Milner, 1983, 1991;
Milner & Smith, 1990; Perzigian et al., 1984), Southeast (Boyd, 1986; Danforth et al., 2007;
Eisenberg, 1991a, 1991b; Hutchinson, 2002, 2004; Lambert, 2000b; Larsen, 2006; Larsen,
Crosby et al., 2002; Larsen & Sering, 2000; Larsen et al., 2007; Parham & Scott, 1980; Rose
et al., 1984), and Northeast (Magennis, 1986; Pfeiffer & Fairgrieve, 1994), as well as a range of
other settings in Mesoamerica and South America (Blom et al., 2005; Cohen et al., 1994;
Hodges, 1989; Hooton, 1940; Hrdlička, 1914; Saul, 1972; Trinkaus, 1977; Ubelaker, 1984,
1992a, 2002; White et al., 1994; Wright, 2006). For some regions where skeletal remains of
foragers (or less intensive agriculturalists) have been compared with agriculturalists, there are
clear temporal increases in porotic lesion prevalence (Cook, 1984; Lallo et al., 1977; Perzigian
et al., 1984; Rose et al., 1984; but see Hodges, 1989). However, skeletal series from large, late
prehistoric Mississippian centers in the American Southeast (Blakely, 1980; Larsen et al.,
2007; Powell, 1988, 1989), contact-era part-time maize agriculturalists in the Great Plains
(Miller, 1995), a large urban center in Mesoamerica (Storey, 1992a), and the coastal desert of
Peru and Chile (Allison, 1984) all display low prevalences of cranial porosities.
The dietary hypothesis presented by El-Najjar does not account for the relatively high
frequencies of porotic lesions in some foraging populations. A number of Pacific coastal
foraging groups with access to iron-rich marine resources have high prevalences of cranial
porosities. Moderate levels of porotic lesions are present in precontact and contact-era
Northwest coast populations (Cybulski, 1977, 1992, 1994, 2006; Keenleyside, 1998, 2006).
In this setting, European-introduced diseases may have prevented adequate iron metabolism
during the contact period (Cybulski, 1994). Presence of porotic lesions prior to contact
indicates that there may have been other important factors, such as blood loss and parasitism
(Cybulski, 1994, 2006).
Late prehistoric foragers from the islands and mainland of the Santa Barbara Channel
Island region of California have higher prevalences than earlier foragers, increasing from
12.8% in the Early period to 32.1% in the Late Middle period (Lambert, 1994; Lambert &
Walker, 1991; Walker, 1986, 2006). Late period populations living on islands located furthest
from the mainland coast have an extraordinarily high prevalence of porotic lesions (73.1% on
San Miguel Island). Walker and Lambert suggest that water contamination explains the
elevated prevalence of the condition. High prevalence of porotic lesions in island populations
coincides with a period of increasing sedentism and population size and a shift from terrestrial
to marine diets. In the Late period, groups became concentrated around a limited number of
water sources. As a result, diarrhea-causing enteric bacteria may have contaminated these
water sources. Ethnographic evidence indicates that island populations preferred eating raw
(versus cooked) fish (Walker, 1986), thus also increasing their chances of acquiring parasitic
infections.
Prevalences of porotic lesions in prehistoric Australian foragers are consistently high in
tropical/subtropical environments and low in desert environments (Webb, 1995). For
example, in southeastern Australia, prevalences range from 62.5% (<21 years) in the Rufus
Valley to 30.0% (<21 years) in the desert. Half of the juvenile crania from the tropics of
northeastern Australia are porotic. Various factors appear to have contributed to iron

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 37

deficiency anemia in Australia, but parasitism is primary. The Murray Valley, southeastern
coast, and tropics provide well-suited environments for support of various intestinal parasitic
organisms, including Trichuris trichuris, Ascaris lumbricoides, Strongyloides stercoralis, and
Enterobius vermicularis. In the tropics, hookworm infection may have been a principal cause.
In living populations occupying the tropics of Australia, some 40% of children are infected
with this helminth. Although it is unknown whether hookworm parasites were present in this
region prior to contact by Europeans (in 1788), had they been, they would have caused the
same types of health problems seen in living groups today.
A limited number of settings in South America show high or temporally increased preva-
lence of porotic hyperostosis and/or cribra orbitalia (Blom et al., 2005; Hrdlička, 1914; Klaus,
Centurion et al., 2010; Klaus & Tam, 2009; Shimada et al., 2004; Ubelaker, 1981, 1992a).
Porotic lesions tend to be low in prevalence in mountainous regions, and appear to be
restricted primarily to late prehistoric and contact-era coastal or near-coastal occupations.
The penchant for coastal settings may reflect a more restricted access by native populations to
fresh, parasite-free water sources in these areas. Ubelaker (1992a) contends that the coastal
pattern of elevated porotic lesion prevalence fits the model of increased anemia due to chronic
helminth disease brought about by population crowding and reduced hygiene. In present-day
Ecuador, hookworm disease is a common and major public health problem in coastal regions.
This distribution in contemporary populations, along with the pattern of prehistoric porotic
lesions, strongly implicates parasitism as a causal factor in northwestern South America.
Moreover, in the contact period of the Lambayeque River valley of the north coast of Peru,
local circumstances during the colonial period, including increased population crowding and
generally deteriorating conditions, provided the necessary conditions leading to reduced
quality of life, including elevated levels of anemia (Klaus & Tam, 2009).
Several investigations suggest that increased prevalence of cranial porosities may be due to
highly localized factors. Lesions have been evaluated in prehistoric and contact-era popula-
tions that inhabited the southeastern US Atlantic coast of Georgia and northern Florida
(Larsen, Crosby et al., 2002; Larsen et al., 2007; Schultz et al., 2001). Maize agriculture –
introduced during the twelfth century – played an important role in changes in health in
native populations, including increased prevalence of nonspecific infections due to popula-
tion aggregation along with other evidence of increased stress (see later). However, compari-
sons of prehistoric foragers and farmers show low prevalences of porotic lesions, 6.5% and
6.2%, respectively. Marine resources contributed significantly to diets in both foragers and
farmers, which may have enhanced iron absorption in these groups. In the contact period,
both the early mission population living on St. Catherines Island, Georgia (Santa Catalina de
Guale; AD 1607 – 1680), and their descendants on Amelia Island, Florida (Santa Catalina de
Guale de Santa Maria; AD 1686 – 1702), have considerably higher prevalences of porotic
lesions, 26.5% and 27.2%, respectively. Archaeological and bone isotope evidence indicates
an increase in maize consumption in the mission groups, and reduction in marine sources of
protein.
The dramatic increase in prevalence of cranial porosities in contact-era coastal Georgia and
Florida populations may be related to similar conditions documented in the Santa Barbara
Channel Islands region. At missions in Spanish Florida, limited and easily contaminated water
sources – wells located next to the settlement – served as primary water sources. Wells in
this subtropical setting are highly susceptible to contamination by parasites and microbes that

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
38 Stress and deprivation during growth and development and adulthood

cause diarrheal infections. During the mission period on St. Catherines Island, a freshwater
stream bordering the mission/village was artificially dammed and also used as a principal
water source (Larsen, Ruff et al., 1992). An abundance of archaeological refuse deposits –
mostly food remains – surrounds and intrudes into this water source today. The accumulation
of refuse during the mission period likely contributed to water contamination, thus also
providing an important source of infections potentially leading to anemia. The general
increase in concentration of population and sedentism during the mission period undoubtedly
fostered poor sanitation and living conditions (Larsen, Crosby et al., 2002). Clearly, anemia is
a likely cause of the elevated prevalence of the lesions. Indeed, histological examination of
lesions reveals morphological characteristics of the diploë consistent with this diagnosis
(Schultz et al., 2001). Moreover, these lesions may have been the result of megaloblastic
anemia caused, at least in part, by deficiencies in vitamin B12 in nursing mothers due to
reduced availability of animal food sources (the only source of vitamin B12) and increased
consumption of maize in the mission setting. Infants born to these mothers have low reserves
of the vitamin. Vitamin B12 is low in the breast milk and gastrointestinal infections and
diarrheal disease associated with poor living conditions would have exacerbated the condi-
tion, increasing the likelihood of anemia.
Data generated in the study of archaeological human skeletons worldwide indicate that the
etiology of cranial porotic lesions can only be understood in relation to multiple stressors.
Although common factors are likely present in many regions (e.g., parasitism, poor diets,
decreased sanitation), these studies also demonstrate that behavioral circumstances unique to
particular settings must be considered when interpreting porotic lesion prevalence. Much
more information is needed on details and circumstances regarding living conditions and
lifestyle (e.g., trash disposal, household and settlement size, dietary practices, food preparation
techniques). Other classes of pathological data need to be considered in understanding health
patterns potentially influencing circumstances resulting in anemia. For example, in at least
one setting, individuals having anemia-related pathology and tuberculosis showed higher
childhood mortality in circumstances where parasitism was minimal (Blom et al., 2005).
Variation in prevalence of cranial porotic lesions in human populations should inform our
understanding of the differential costs of disease stress in past societies. In some settings,
there is a consistently higher prevalence of porotic lesions in adult women than in adult men
(Dickel, 1991; Webb, 1995). For Australia, Webb (1995) suggests that a higher prevalence in
women than in men reflects the stresses of “[c]hildbearing, lactation, menstruation and the
imposition of food taboos.” Given that the pathological condition largely reflects childhood
anemia, it seems unlikely that parity status, lactation, or menstruation can explain variability
in porotic lesion prevalences. These differences suggest, however, that the growth period may
have involved greater anemia stresses in female children than in male children.
Comparison of porotic lesion prevalence across social ranks in prehistoric stratified soci-
eties suggests some differences in iron deficiency anemia. Elite individuals in several settings
in North America and South America have a lower prevalence of porotic lesions than non-
elite individuals. This difference has been reported at prehistoric Mississippian localities from
Moundville, Alabama (2.5%, elite; 9.9%, non-elite) and at Toqua, Tennessee (5%, mound
burial; 21%, village burial) (Parham & Scott, 1980; Powell, 1988, 1992a). At Mound 72 in the
late prehistoric Cahokia site, Illinois, high-status individuals lacked lesions, whereas 12.5% of
low-status individuals – female sacrificial victims – have the condition (Fowler et al., 1999).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 39

Similarly, higher-status individuals (especially women) have lower prevalence than lower-
status people documented in the Middle Sicán burials in late prehistoric Peru (Shimada et al.,
2004), and most elite individuals at Copan, Honduras, have no lesions, whereas most of the
non-elite have lesions (Storey, 1998). Overall, porotic hyperostosis in this setting is nearly five
times greater in non-elite than in elite individuals. These differences suggest, therefore, that
high-status individuals may have been buffered against stressors that result in the patho-
logical condition, consistent with the notion that political and socioeconomic inequality is
linked to stress and quality of life.
In summary, a consensus has emerged that porotic hyperostosis and cribra orbitalia in
archaeological skeletal samples are due to anemia in the vast majority of cases. Genetic
anemias may have contributed to elevated porotic lesion prevalence in the past, but it is
unlikely that they would have occurred in appreciable frequencies. Moreover, significant
presence of porotic lesions in areas of the world where genetic anemias (e.g., sickle cell
anemia, thalassemia) did not occur prior to contact by Europeans – such as Australia and
the Americas – can be explained only by negative environmental factors. Moreover, only
marrow hypertrophy activated by overproduction of red blood cells, a condition associated
with megaloblastic or hemolytic anemias, will cause porotic hyperostosis and cribra orbitalia,
not iron-deficiency anemia.
Angel argued that the etiology for porotic hyperostosis and cribra orbitalia is the same,
resulting in him conflating the two terms into porotic hyperostosis (and see Blom et al., 2005;
Stuart-Macadam, 1989). However, it is becoming increasingly clear that while porotic hyper-
ostosis affecting the flat bones of the cranial vault is generally caused by anemia and is
associated with hypertrophy of the cranial vault marrow, there are a large number of cases
where cribra orbitalia is associated with subperiosteal inflammation and not marrow hyper-
trophy. For example, Wapler and coworkers (2004) found that 26% of cribra orbitalia is
associated with inflammation and not marrow expansion. That is, other pathological pro-
cesses cause circumstances resulting in inflammation, such as vitamin C and D deficiencies
(scurvy and rickets, respectively), and trauma-related hematomas. The hematomas become
plaques of new, vascular subperiosteal bone that superficially look like cribra orbitalia
associated with marrow hypertrophy (Brickley & Ives, 2006, 2008; Schultz, 2001; Sloan
et al., 1999; Walker et al., 2009). This process is common in childhood, a period of life when
the periosteum is not well attached (Ma’luf et al., 2002). These non-anemia lesions of the eye
orbits are associated with scurvy (Brickley & Ives, 2008; Halcrow et al., 2014; Klaus, 2014b;
Ortner et al., 1999), owing to the fact that vitamin C (ascorbic acid) deficiency renders the eye
orbit prone to minor trauma due to normal movement of the ocular muscles. As a result, there
is a detachment of the periosteum in the superior orbital roof and associated bleeding and
formation of new, highly vascularized bone. Similar kinds of porosities have been well
documented in other cranial areas such as the sphenoid, alveolar bone, and palatine processes
where blood vessels are close to skin surfaces and are subject to greater bleeding, especially in
association with scorbutic episodes (Figure 2.9a, b) (Geber & Murphy, 2012; Mays, 2008;
Klaus, 2014b; Ortner et al., 2001). These porotic lesions and bony plaques also co-occur on
long bones, and like the eye orbits, are bilateral (Lovász et al., 2014).
Cribra orbitalia and porotic hyperostosis lesion formation are highly age-specific
(Figure 2.10). Although the two conditions are present in older juveniles and adults in
archaeological remains, most individuals with active, unhealed lesions are young juveniles

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
40 Stress and deprivation during growth and development and adulthood

(a)

(b)

Figure 2.9 Abnormal bone growth and porosity in glabellar region (a) and on greater wing
of sphenoid (b), diagnostic of scurvy; Chornancap, Peru. (Klaus, 2014b; reproduced with
permission of author and Elsevier Ltd.) (A black and white version of this figure will appear
in some formats. For the color version, please refer to the plate section.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 41

45

40 Cribra Orbitalia

Porotic Hyperostosis
35

30
% Showing Condition

25

20

15

10

0
0 10 20 30 40 50 60 70
Age (years)
Figure 2.10 Age distribution of cribra orbitalia and porotic hyperostosis, from the History of
Health in the Western Hemisphere database (Steckel & Rose, 2002). (Data from Walker
et al., 2009; reproduced with permission of John Wiley & Sons, Inc.)

(less than five years of age) for both conditions, regardless of geographic or cultural circum-
stances (Blom et al., 2005; Fairgrieve & Molto, 2000; Lallo et al., 1977; Larsen, Ruff et al.,
1992; Mensforth et al., 1978; Milner & Smith, 1990; Miritoiu, 1992; Mittler & Van Gerven,
1994; Ribot & Roberts, 1996; Salvadei et al., 2001; Stodder & Martin, 1992; Walker, 1986;
Webb, 1995; Wright, 2006; and many others). This age pattern indicates that the lesions form
during childhood (Stuart-Macadam, 1985; Walker et al., 2009). With regard to porotic
hyperostosis in particular, red blood cell production occurs in the cranial marrow in juveniles
but not the yellow marrow that characterizes adolescents and adults.
There is considerable variation of orbital versus non-orbital (cranial vault) lesions across
human populations. In prehistoric Ecuador, most lesions are associated with the cranial vault
(Ubelaker, 1992a), but in native Australians and Arctic populations, the lesions are associated
largely with the eye orbits (Dabbs, 2011; Webb, 1995). Some argue that cribra orbitalia is an
earlier manifestation of anemia than porotic hyperostosis (Blom et al., 2005; Carlson et al.,
1974; Lallo et al., 1977). Walker and coworkers (2009; Wapler et al., 2004) make the case that
at least some cribra orbitalia have little to do with iron-deficiency anemia and that common
etiology with porotic hyperostosis is not necessarily the case.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
42 Stress and deprivation during growth and development and adulthood

Figure 2.11 Harris lines on juvenile tibia (left) and femur (right); anatomical specimens. The
dashed lines indicate the contours of the growth disruption. (From Garn et al., 1968;
reproduced with permission of Eastman-Kodak Company.)

2.5.2 Skeletal growth (Harris) lines


On many skeletal elements, including long bones and round or irregular skeletal elements
(e.g., scapula, ischium, ilium), radiopaque lines may be visible in X-rays that follow growth
contours, “topographically mapping the history of the bone” (Garn et al., 1968:58)
(Figure 2.11). Lines range in thickness from less than one millimeter to more than a centi-
meter, and are thickest in areas of rapid growth, such as the distal tibia and femur (Dreizen
et al., 1964; Garn et al., 1968).
Although originally considered symptomatic of rickets (Wegner, 1874), studies of living
populations and animal studies link transverse lines to many other conditions potentially
resulting in metabolic insult, including dietary insufficiencies (Blanco et al., 1974; Dreizen
et al., 1956, 1964; Garn et al., 1968; Harris, 1931, 1933; Martin et al., 1985; Park & Richter,
1953; Platt & Stewart, 1962; Stewart & Platt, 1958), disease (Acheson, 1959; Harris, 1931,
1933), trauma from minor surgery and immunization (Garn et al., 1968), fracture (Ferrozo
et al., 1990), lead poisoning (Caffey, 1931), and the physiological and psychological impact
of weaning (Clarke & Gindhart, 1981). Most lines appear to form after six months of life,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 43

peaking sometime during the first five years (Clarke, 1980; Clarke & Gindhart, 1981; Dreizen
et al., 1964).
These lines are commonly referred to as growth arrest lines or growth recovery lines
representing the recovery phase following growth arrest (Garn et al., 1968). The link with
stress episodes, however, remains unclear as a number of studies have shown no association
between Harris lines and illness (Papageorgopoulou et al., 2011), and experimental studies
show that more lines are associated with rapid growth in the absence of nutritional stress
(Alfonso-Durruty, 2011). When the epiphysis commences growth following the stress event,
mineralization at the growth plate continues in the absence of epiphyseal cartilage deposition.
The bilateral nature of transverse lines and association with growth zones (e.g., same location
in distal left and right tibiae: Garn & Baby, 1969; Grolleau-Raoux et al., 1997; McHenry,
1968) provide a strong argument for their systemic origin.
Another difficulty of using Harris lines for assessing stress is the high degree of frequency
variation in relation to individual health history. The study of living individuals with known
stress histories reveals the presence of numerous lines in clinically normal children (Garn
et al., 1968), and few lines in children who are well below weight-for-age (Walimbe &
Gambhir, 1994). These findings and the lack of a close association between transverse lines
and disease episodes in archaeological populations (Mensforth, 1981) and in living popula-
tions (Marshall, 1968) suggest that this stress indicator should be interpreted cautiously in
bioarchaeological analysis, especially when considering health status and its relationship to
specific behavioral, environmental, and dietary adaptations.
Analyses of transverse lines in archaeological remains are largely built on the assumption
that they represent stress events. Comparisons of transverse lines in early and late prehistoric
foragers in central California show a general decrease in frequency, possibly indicating
improved reliability of food sources (Dickel et al., 1984; McHenry, 1968). This argument runs
counter to an abundance of evidence drawn from the study of other pathological conditions
and stature showing an increase in stress in populations in the Central Valley (Ivanhoe, 1995)
and to the south in the Santa Barbara Channel Islands region (Lambert, 1993).
Similarly, other prehistoric foragers show variable frequencies in transverse lines. High
latitude Arctic populations express elevated prevalences (>30%–50%: Buikstra, 1976b; Lob-
dell, 1984, 1988; Steffian & Simon, 1994). In several of these settings where even spacing of
lines has been observed, regularity or periodicity of stress – perhaps on a seasonal basis – is
inferred (Buikstra, 1976b; Lobdell, 1988). The common occurrence of transverse lines in these
populations indicates that metabolic stress is more characteristic of Arctic lifeways than has
often been assumed (Steffian & Simon, 1994). In addition to nutritional deficits, other
stressors associated with this setting include the constant variation in the ratio of light to
dark, placing significant demands on the body, especially in children (Condon, 1983). Sea-
sonal or other changes in the intensity and duration of daylight lead to various changes in
health and mental functioning, especially in circumstances involving depletion of light. The
study of living Inuit in the central Canadian Arctic indicates that January is a peak time of
disease susceptibility, primarily because of extremely low temperatures, low ambient humid-
ity within and outside dwellings, and lowered sunlight (Condon, 1983). This may be exacer-
bated by the effects of desynchronization of natural physiological rhythms and lack of sleep.
Temporal comparisons of transverse lines in archaeological samples presents a mixed
picture, and sometimes contrary to expectation based on other stress indicators.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
44 Stress and deprivation during growth and development and adulthood

Comparisons of three successive periods at Dickson Mounds – Late Woodland (AD 950 –
1050), Mississippian Acculturated Late Woodland (AD 1050 – 1200), and Middle Mississippian
(AD 1200 – 1300) – reveal a decrease in tibial transverse lines (Goodman & Clark, 1981).
Similarly, lines decreased in frequency in the comparison of foragers and farmers in the Ohio
River valley (Cassidy, 1984; Perzigian et al., 1984) and in the Caddo region of the southeastern
United States (adjoining the states of Texas, Oklahoma, Arkansas, Louisiana) (Rose et al.,
1984). These trends suggesting a decrease in stress are puzzling because most other indicators
of morbidity show a highly consistent pattern of increase in stress and reduced health status
(e.g., enamel defects, nonspecific infection).
The use of Harris lines for documenting stress in past populations is clouded further by the
fact that lines have a tendency to fade or vanish with advancing age due to bone remodeling.
In a study of living populations comprised of individuals of known stress history, lines
showed a decrease in width with advancing age; some lines disappeared while others were
inexplicably retained well into adulthood (Garn & Schwager, 1967; Garn et al., 1968). This
very mixed record suggests that Harris lines may not reflect stress histories, at least with any
degree of clarity. Rather, these lines correspond with normal growth and normal growth spurts
and should not be regarded as a stress indicator (Alfonso-Durruty, 2011; Papageorgopoulou
et al., 2011).

2.5.3 Growth disruption of dental tissues: enamel defects


Growth of tooth enamel commences at the incisal or cuspal terminus of the crown and
proceeds in a uniform fashion to completion (Figure 2.12). The enamel is first laid down by
ameloblasts (enamel-producing cells) secreting a highly proteinaceous, mostly nonmineral-
ized, matrix. This matrix then mineralizes into an acellular material composed mostly (~97%)
of inorganic salts, thus forming the fully mature enamel (Goodman & Rose, 1990; Nanci,
2013). The enamel matrix is deposited in a series of structural increments demarcated by striae
(or lines or bands) of Retzius. Like skeletal tissue, the formation of enamel is a regular process
that is subject to factors that may either slow or stop growth. Ameloblasts are especially
sensitive to metabolic insults arising from nutritional deficiencies or disease, or both. Because
enamel does not remodel and it preserves better than any other hard tissue, developmental
disturbances provide an excellent source of information toward reconstructing a retrospective
history of stress and morbidity of human populations, past or present.

Macrodefects
Contrary to Harris lines, there is a clear link between enamel defects and stress. Indeed,
virtually any environmental factor leading to metabolic disturbance will result in visible
changes in the structure of enamel. Ameloblasts are especially sensitive to even minor
physiological disruptions. Enamel defects arising from physiological perturbation have been
documented most frequently as visible alterations of the tooth surface, especially hypoplasias,
and to a lesser extent, hypocalcifications. Hypoplasias are measurable defects characterized as
deficiencies in the amount or thickness of enamel (Goodman & Rose, 1990; Hillson, 2008,
2014; Seow, 1991; Suckling, 1989). They vary in appearance from small pits or furrows to
large, deep grooves or even large areas of missing enamel. Typically, these defects are
horizontal grooves that are called chronologic or linear enamel hypoplasias (Figure 2.13).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 45

(a) (b)
Enamel

Perikymata

Striae of
Retzius
(c)
Dentin
Occlusal border
of defect

Occlusal wall
perikymata
(stress episode)
Cervical/occlusal wall junction
cervical wall perikymata
Pulp (recovery period )
Cervical border
of defect

Figure 2.12 Cross-section of tooth displaying the relationship between striae of Retzius (a) and
perikymata (b). Episodes of stress disrupt enamel growth, which are followed by recovery periods (c).
The resultant enamel surface captures this period when little or no enamel matrix was formed in
the tooth. (From Blatt, 2013; Fejerskov & Thylstrup, 1986; Guatelli-Steinberg et al., 2004; and Hubbard
et al. 2009; reproduced with permission of authors, John Wiley & Sons, Inc., and Munksgaard
International Publishers.) (A black and white version of this figure will appear in some formats. For the
color version, please refer to the plate section.)

Figure 2.13 Maxillary dentition showing enamel hypoplasia on incompletely erupted


central incisors; anatomical specimen. (From Larsen, 1994; photograph by Barry Stark;
reproduced with permission of John Wiley & Sons, Inc.) (A black and white version of this
figure will appear in some formats. For the color version, please refer to the plate section.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
46 Stress and deprivation during growth and development and adulthood

Circular defects of thinned or missing enamel are also well documented, especially in
deciduous teeth (Halcrow & Tayles, 2008), but the linear form is the most common type of
hypoplasia. Regardless of form – linear or circular – the color and hardness of hypoplastic
enamel is normal.
Hypocalcifications are enamel defects involving change in color or opacity reflecting
variation in enamel quality or hardness. The enamel surface is usually smooth and appears
intact. Hypoplasias occur when ameloblasts fail to produce the normal thickness of enamel
matrix during enamel development, whereas hypocalcifications appear to result from a
disruption of the mineralization process during the maturation stage of enamel development.
However, this dichotomy may not be so clear cut as hypocalcifications have been experi-
mentally documented in the initial stage of enamel formation (Suckling, 1989).
Hypoplasias result from three potential causes, including hereditary anomalies, localized
traumas, and systemic metabolic stress, such as from illness, disease, and malnutrition (Goodman &
Rose, 1991; Hillson, 1996). Defects arising as hereditary anomalies or as localized traumas are
rare in human populations, indicating that the vast majority of hypoplasias seen in contemporary
and archaeological populations are linked with systemic physiological stress (Hillson, 2008;
Hillson & Bond, 1997; King et al., 2005). The causal stressors associated with hypoplasias are
numerous and varied. Clinical and epidemiological investigations in living populations docu-
ment associations with systemic diseases, neonatal disturbances, and nutritional deprivation
(reviewed in Hillson, 1996; Pindborg, 1982). Experimentally induced stress in laboratory animals
have also shown the direct link between enamel deficiency and stress (Kreshover, 1944;
Kreshover & Clough, 1953a, 1953b; Suckling & Thurley, 1984; Suckling et al., 1983, 1986).
Studies of non-human primates with known life histories reveal links between enamel defects,
life events, and local ecology (Dirks et al., 2002). Thus, enamel defects are a nonspecific indicator
of physiological stress (Goodman & Rose, 1990; Kreshover, 1960; Pindborg, 1982).
Ecological factors are critical for understanding the prevalence and pattern of enamel
defects in human populations. Studies of living populations with dietary deficiencies show
the primacy of nutrition in the development of normal enamel. Analysis of individuals born
during the starvation famine of 1959 – 1961 in the People’s Republic of China reveals that
enamel forming during the famine is highly defective, unlike the enamel that formed either
before or after the famine (Zhou & Corruccini, 1998). For the population as a whole, there was
a clear increase in stress as is clearly revealed in the spike in enamel defects during the famine
period (Figure 2.14). Rural individuals have more defects than urban individuals, a pattern
consistent with records documenting more stress in the rural than in the urban population.
Similarly, forced resettlement, declining nutritional quality, and poor living conditions gen-
erally of Warlpiri people in Northern Territory, Australia, and the Tupí-Mondé speakers in
Brazil in the twentieth century resulted in a profound increase in stress and a correspondingly
marked increase in prevalence of hypoplasias (Littleton, 2005; Santos & Coimbra, 1999).
Enamel hypoplasias show a predilection for anterior teeth and for the cervical and middle
thirds of tooth crowns (Barrett & Blakey, 2011; Condon & Rose, 1992; Goodman & Arme-
lagos, 1985a, 1985b; Hutchinson & Larsen, 1988, 2001; King et al., 2005; Li et al., 1995;
Pedersen & Scott, 1951; Starling & Stock, 2007; Wright, 1997; Zhou & Corruccini, 1998).
This is the case because of the layering nature of enamel deposition in forming the cuspal
component of the tooth crown (Figure 2.12). That is, enamel grows in a series of dome-like,
incremental layers, building one on top of the other, eventually forming the crown cusp(s)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 47

70
Whole sample
Rural
60
Prevalence of individuals with LEH (%) Urban

50

40

30

20

10

0
Pre-famine Famine Post-famine
Figure 2.14 Comparison of linear hypoplasia (LEH) prevalence (percentage of affected
individuals) among living rural (Qingji Township) and urban populations (Shanghai City) in
People’s Republic of China. (Adapted from Zhou & Corruccini, 1998; reproduced with
permission of authors and John Wiley & Sons, Inc.)

and occlusal surface of the tooth. Following completion of the cuspal enamel, the sides of
the tooth crown form, also in a series of discrete layers comprising the imbricational enamel
(FitzGerald & Rose, 2008; Guatelli-Steinberg, 2008; Hillson, 1996, 2008, 2014; Hillson &
Bond, 1997). The layers of cuspal and imbricational enamel are separated by incremental
lines representing the terminus of each depositional episode. These lines, or striae of Retzius,
appear on the surface of the tooth as perikymata.
The top striae of Retzius forming the cusp of the tooth are not exposed on the surface of the
tooth. Thus, except for the final dome of cuspal enamel, all earlier-formed cuspal enamel (20%
– 50% of the crown) is hidden, as are any associated enamel defects (FitzGerald & Rose, 2008;
Hillson, 2008, 2014; Reid & Dean, 2000). The perikymata along the crown sides are, however,
exposed. When insufficiencies of enamel deposition occur, it is in the imbricational enamel
where enamel defects manifest in a visible fashion. The occlusal wall of the hypoplasia
represents the period of time when the stress event occurred (Guatelli-Steinberg, 2008; Hillson
& Bond, 1997). The cervical wall of the enamel defect represents the period of recovery. Thus,
together the occlusal and cervical walls form the hypoplasia (Figure 2.12).
The record of hypoplasia defines a relationship between poor living circumstances and
increased likelihood of increased stress and poor health (see Chapter 3). Few analyses of this
compelling record address the relationship between stress and social hierarchy or inequality.
In a pioneering test case investigating the relationship between social organization and
human biology, analysis of hypoplasia reveals a striking difference in the stress experience

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
48 Stress and deprivation during growth and development and adulthood

between elite and non-elite Middle Sicán society in pre-contact Peru (c. AD 1000). That is,
hypoplasia is seven times more common in the non-elite than in the elite in this setting (Klaus,
2014b). This finding is consistent with other settings (Goodman, Armelagos et al., 1984;
Swärdstedt, 1966; and see Goodman, 1998; Goodman and Martin, 2002).

Microdefects
Histological structures in dental enamel known as Wilson bands (or accentuated or patho-
logical striae of Retzius) provide a detailed record of growth disruption (FitzGerald & Rose,
2008; Goodman & Rose, 1990; Simpson, 1999, 2001). Wilson bands are thin layers of
abnormally structured enamel marking the position of the active ameloblasts at the time of
the insult (Figure 2.15). A highly common association between these incremental structures
and life history is the transition from the intra- to extra-uterine environment, resulting in a
distinctive “neonatal line” on the forming teeth, namely in the deciduous teeth and the
permanent first molars as these teeth begin to form in utero (Antoine et al., 2009; Hillson,
2014; Schour, 1936; Schwartz et al., 2010; Whittaker & Richards, 1978). Wide Wilson bands
are associated with traumatic births and perinatal strain (Eli et al., 1989), suggesting that the
width of Wilson bands may represent an indicator of stress severity. Importantly, bands that
match within a single dentition demonstrate the presence of systemic stress for the person

Figure 2.15 SEM micrograph ( 230) of a polished and acid-etched longitudinal section
from a maxillary central incisor, showing a Wilson band (arrows); Santa Catalina de Guale,
Amelia Island, Florida. (Photograph by Scott W. Simpson.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 49

represented (FitzGerald et al., 2006). Analysis in the prehistoric-colonial era series from
Spanish Florida reveals a clear association between Wilson bands and tooth size (Simpson,
1999). Individuals with at least one Wilson band have smaller tooth size (mandibular canine
cervical breadth) than individuals without Wilson bands. This finding is consistent with the
notion that stressed individuals, including short-term stress, are likely less able to resist
physiological stress and not able to achieve full growth potential.
Wilson bands are not always associated with a surface macrodefect such as a hypoplasia
(Bullion, 1986; Condon, 1981; Condon & Rose, 1992; Goodman & Rose, 1990; Hillson, 2008;
Rose, 1977; Wright, 1990). The lack of a consistent association between macro- and micro-
defects suggests that their stress etiologies may be different. Wilson bands appear to represent
brief periods of stress lasting for hours or days, whereas hypoplasias appear to represent long-
term stress lasting from weeks to months.

n = 30 / n = 39 n = 22 / n = 16 n = 22 / n = 19

4.8/5.3 4.2/5.0
4.8/5.1 3.8/4.4
4.4/4.8
4.4/4.6
4.0/4.3 3.4/3.9
3.9/4.1
3.5/3.8 2.9/3.4
3.5/3.7
3.1/3.4 2.5/2.9
3.2/3.3
2.8/3.0 2.9/2.9 2.2/2.4
2.6/2.7 1.9/2.0
2.5/2.7
2.4/2.4
2.3/2.4 1.7/1.8
2.2/2.2
2.1/2.2 1.5/1.6
2.0/2.0
1.9/1.9 1.3/1.3
1.8/1.8
1.7/1.7 1.4/1.5 1.1/1.1
1.6/1.7
1.0/1.0 0.8/1.0
1.9/2.0 1.2/1.1 1.0/1.1
1.3/1.3 1.1/1.3
2.1/2.3
1.5/1.5 1.3/1.5
2.4/2.7
1.7/1.8 1.5/1.7
2.8/3.1
1.9/2.1 1.7/2.0

3.2/3.6 2.2/2.4 2.0/2.3


2.6/2.8 2.3/2.6
3.7/4.2
2.7/3.0
3.0/3.3
4.2/4.9 3.1/3.4
3.4/3.7
4.7/5.6 3.4/3.8
3.8/4.2
5.2/6.2

n = 25 / n = 13 n = 27 / n = 13 n = 29 / n = 15

Figure 2.16 Comparison of mean chronological ages associated with enamel formation in
anterior maxillary and mandibular canines. Ages have been divided according to African
(left) and European (right) samples. (Reid & Dean, 2006; reproduced with permission of
Elsevier Ltd.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
50 Stress and deprivation during growth and development and adulthood

40

Stillwater
35
Prevalence (%) of enamel hypoplasias
Georgia Agricultural

30 Georgia Preagricultural

25

20

15

10

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5
Age (years)
Figure 2.17 Stillwater maxillary permanent central incisor hypoplasia frequencies
(percentage per half-year age group) compared with Georgia coast foragers and farmers.
(Adapted from Hutchinson & Larsen, 1995; reproduced with permission of the American
Museum of Natural History.)

Stress chronology
Because metabolic insults leading to growth disruption affect only the part of the tooth that is
in the process of forming, location of the disturbance on the tooth crown provides a precise
chronologic indicator of stress history. Tooth enamel begins to form at about four months in
utero, beginning with the deciduous first incisors, and is complete when the crowns of the
permanent third molars are fully formed at about 12 years of age (Nanci, 2013; Smith, 1991).
The location of the position of an enamel defect (e.g., hypoplasia) relative to the cemento-
enamel junction can be used to plot the age of disturbance (Reid and Dean, 2006) (Figure 2.16).
Earlier researchers suggested that there is a preprogramed stress clock in humans (Massler
et al., 1941; Sarnat & Schour, 1942). However, the body of evidence building over the last five
decades indicates little support for a universal model of timing of growth disruption (Hillson,
2008). Rather, growth disruption is due to individual and local circumstances involving
environmental (or other) stresses. That is, even within relatively circumscribed cultural and
temporal settings, specific skeletal series show very different patterns and frequency of
enamel defects (Hutchinson & Larsen, 2001; King et al., 2005; Lieverse, Link et al., 2007;
Temple, 2007).
Although the relationship between enamel defects and age has been recognized since the
nineteenth century (Talbot, 1898), this chronological approach has been applied to archaeo-
logical remains only recently. The pioneering investigation of Medieval-era dentitions from

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 51

Westerhus, Sweden, by Swärdstedt (1966) revealed that hypoplasias peaked in the two-to-
four-year period, a pattern that has been identified in many other archaeological samples
(Corruccini et al., 1985; Goodman et al., 1980; Hillson, 1979; Hodges, 1989; Hutchinson &
Larsen, 1995; Martin et al., 1991; Powell, 1988; Storey, 1992a, 1992b) (Figure 2.17).
The tendency for hypoplasias to occur after the first year or so of life as documented in
archaeological samples suggests that stress may be linked to the negative effects of weaning
(Coppa et al., 1995; Corruccini et al., 1985; Lanphear, 1990; Lillie, 1996; Moggi-Cecchi et al.,
1994; Ogilvie et al., 1989; Simpson et al., 1990; Ubelaker, 1992b; Webb, 1995; and many
others). A test of the weaning hypothesis, based on the study of historical records and
archaeological dentitions from enslaved African-descent populations living in the mid-
Atlantic region of the United States (Maryland and Virginia), reveals that the peak frequencies
of hypoplasias are in the 1.5 – 4.5-year age intervals, whereas weaning took place only nine
months to one year after birth (Blakey et al., 1994). This discrepancy between age pattern of
hypoplasias and weaning led Blakey and coworkers (1994; and see Barrett & Blakey, 2011) to
conclude that weaning was not the primary causal factor leading to enamel defects. That is,
other stresses of enslavement, including nutritional problems, poor hygiene, and illness, were
likely responsible for the age pattern of physiological perturbation in this setting. Thus,
weaning as a cause for the age profile of enamel defects in archaeological settings, usually
between two and four years, is not the sole factor leading to growth arrest as documented by
enamel hypoplasias (and see Hassett, 2014; Katzenberg et al., 1996; Saunders & Keenleyside,
1999). Rather, the problems identified in African-descent populations from the mid-Atlantic
region reflect a period of weakened immunity in early life and reliance on foods that do not
include breast milk. It is this set of circumstances, of which weaning is one, that give the
infant a poor growth environment and result in growth arrest.
Insights into the timing of short-term stress from Wilson bands provide a similar conclu-
sion. Analysis of timing of Wilson bands in the first year of postnatal life in a large series of
deciduous dentitions from first- to third-century AD Rome at the cemetery of Isola Sacra
reveals peak prevalence in the second through fifth months and the sixth through nine
months. This age pattern of short-term growth arrest closely matches with historical sources
indicating two periods of infant frailty, the first occurring around three months when the
infant is first exposed to non-milk foods (e.g., bread softened with milk or wine) and a second
when the infant begins to eat solid foods (FitzGerald et al., 2006). This supports the weaning
hypothesis, but almost certainly, this setting also had other negative circumstances that
contributed to poor growth during these critical months of life.

Duration/severity of stress
Because enamel is deposited in consecutive layers from the incisal/occlusal surface to the
cementoenamel junction over some interval of time, hypoplasia width should represent a
quantification of duration of stress events (Blakey & Armelagos, 1985; Ensor & Irish, 1995;
Hassett, 2014; Hutchinson & Larsen, 1988, 1990, 1995, 2001). Based on their experimental
work with laboratory sheep, Suckling and coworkers (1986; and see Suckling, 1989) con-
cluded that the severity of the stress insult determines the magnitude of the hypoplasia. Thus,
hypoplasia size may reflect either stress duration or severity, or perhaps some unknown
combination of both.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
52 Stress and deprivation during growth and development and adulthood

On the other hand, hypoplasia size per se is not the best estimate of duration of the stress
episode, largely because enamel does not grow at a constant, linear rate (Guatelli-Steinberg,
2008; Hassett, 2014; Hillson, 2008; Hubbard et al., 2009; Reid and Dean, 2006). Simply,
enamel rates vary according to location in the tooth crown. Alternatively and far more
precisely, Hillson and Bond (1997) make the case that the duration of the episode of
developmental disruption is best estimated by counting the number of perikymata within
the margins of the defect, especially with regard to the number of perikymata in the defect’s
occlusal wall, as this is the actual period of growth disruption (Hillson & Bond, 1997)
(Figure 2.12). For example, a chronic disruption may have more than 20 perikymata, whereas
a shorter-term disruption may have fewer than five perikymata. In a Jomon series (Aichi,
Japan), Temple and coworkers (2012) documented a range from one to 20 perikymata in a
single individual, representing stress duration from 12 days to 130 days. All stress events took
place between the ages of 1.2 and 3.5 years. Precise estimates of duration of the stress episode
may be determined as each stria represents about seven to nine days of growth (FitzGerald,
1998; Guatelli-Steinberg et al., 2005; Reid & Dean, 2006). For example, an occlusal wall of a
defect containing nine perikymata represents approximately three months of growth disrup-
tion. Thus, not only do these counts provide precision with regard to timing of stress events,
they also provide an important measure of stress episode duration and inter- and intra-
population variation (Hubbard et al., 2009; Temple et al., 2012, 2013).

Stress histories in human populations


The connection between poor living conditions and enamel defect prevalence is well
supported by epidemiological studies of contemporary human populations. In general,
individuals from developed nations tend to have far lower hypoplasia prevalences than
individuals from underdeveloped nations (Goodman & Rose, 1991). In this regard, less than
10% of individuals from developed nations have one or more hypoplasias, whereas hypo-
plasias are commonplace in many underdeveloped settings or in disadvantaged subgroups
of populations with poorer diets, more disease, or some combination of undernutrition and
disease (Anderson & Stevenson, 1930; Baume & Meyer, 1966; Dobney & Goodman, 1991;
Enwonwu, 1973; Goodman et al., 1987, 1991, 1992; Infante, 1974; Infante & Gillespie,
1974, 1977; Li et al., 1995; Lukacs & Joshi, 1992; Massler et al., 1941; May et al., 1993;
Pedersen & Scott, 1951; Rugg-Gunn et al., 2000; Sawyer & Nwoku, 1985; Sweeney et al.,
1971; Zhou & Corruccini, 1998).
Several case studies are especially informative regarding the link between stress, socio-
economic status, and life history. Children from villages in the Solis Valley of the Temascal-
cingo region of the Mexican highlands display prevalences of hypoplasias that document the
relationship between poor growth status and physiological disruption (Goodman et al., 1992).
Children with enamel defects have reduced body weights and heights-for-age in comparison
with children who lack defects. Predictably, children with hypoplasias tend to be from families
of lower socioeconomic status living under conditions of malnutrition and poor sanitation.
The association between negative environments and defect prevalence is well illustrated in
settings involving selected dietary supplementation (Dobney & Goodman, 1991; Goodman
et al., 1991; May et al., 1993). In rural Mexico and Guatemala, for example, children receiving
dietary supplements have far fewer linear enamel hypoplasias than their non-supplemented
peers. In Guatemala, children who were ill more than 3.6% of the time had more hypoplasias

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 53

than other children (May et al. 1993). Comparable patterns of morbidity differences are
reported for Wilson band prevalence in living populations. Children with histories of chronic
systemic disease in Sheffield, England, have elevated prevalence of Wilson bands in their
deciduous teeth (Hillier & Craig, 1992; and compare with FitzGerald et al., 2006, for an
archaeological setting).
The study of dentitions from paleontological and archaeological contexts adds a great deal
to our present understanding of the history of human stress and its complex links with
environment, culture, and biology. Unlike most stress indicators, enamel defects have been
systematically investigated in a range of early hominins, including australopithecines (Tobias,
1967; White, 1978), early archaic Homo sapiens (Bermúdez de Castro & Pérez, 1995), and
Neanderthals (Guatelli-Steinberg, 2008; Guatelli-Steinberg et al., 2004; Hutchinson et al.,
1997; Molnar & Molnar, 1985a; Ogilvie et al., 1989). Early archaic H. sapiens from Atapuerca,
Spain (Bermúdez de Castro & Pérez, 1995), and Neanderthals (Ogilvie et al., 1989), possess low
to moderate prevalences of hypoplasias (permanent dentition: 12.8%, Atapuerca; 41.9%,
Krapina). Each of the two series displays similar age patterns of hypoplasias, including two
primary peaks, the first in early childhood and a second in late childhood. The late peak is
especially interesting because it represents stress events affecting posterior teeth, teeth which
rarely exhibit enamel defects in modern human populations (Ogilvie et al., 1989). Thus, the
earlier peak may reflect early childhood stress (e.g., weaning), and the later peak may
represent overall high levels of systemic stress. Infection may have been a cause of growth
disruption, but the Atapuerca and Krapina samples show very low prevalences of skeletal
infection. Instead, nutritional deficiencies – perhaps during periodic food shortages – were the
most likely causative factor (Guatelli-Steinberg et al., 2004; Hutchinson et al., 1994; Ogilvie
et al., 1989). The record suggests overall that developmental stress levels in Neanderthals by
various measures are high, including elevated prevalence of dental enamel defects and
fluctuating asymmetry (Barrett et al., 2012), but not remarkably so in comparison with
modern foragers from archaeological contexts (Guatelli-Steinberg et al., 2004; Hutchinson
et al., 1994).
Temporal comparisons of enamel defect prevalence in Holocene populations undergoing
dietary and behavioral changes show clear trends in physiological stress. Especially striking
are changes observed in human populations undergoing the shift from foraging to agriculture
or agricultural intensification and increased population density. In general, these comparisons
reveal increases in prevalence of enamel defects, especially in populations inhabiting the
Eastern Woodlands of North America (Cassidy, 1984; Cook, 1984; Goodman, Lallo et al.,
1984; Perzigian et al., 1984; Rose et al., 1984), Latin America (Márquez Morfín and Storey,
2007; Ubelaker, 1984, 1992b; but see Hodges, 1989; Lukacs & Joshi, 1992, for exceptions),
and to a lesser extent, in Asia (Rathbun, 1984; Smith, Bar-Yosef et al., 1984; but see Temple,
2010; Temple & Larsen, 2013; Yamamoto, 1992) and Africa (Starling & Stock, 2007). This is
not a universal trend, however. In comparison with late prehistoric and contact-era maize
agriculturalists from one setting in the American Southeast, there is a decrease in hypoplasia
prevalence (Hutchinson & Larsen, 1988; Larsen, Crosby et al., 2002; Larsen & Hutchinson,
1992; Larsen et al., 2007; Simpson et al., 1990). Hypoplasia widths increase in this setting.
These trends reflect a decline in the number of individuals affected, but stress episodes were of
either longer duration, greater severity, or both. Other late prehistorics or contact-period
agriculturalists in the New World express generally high frequencies of hypoplasias, which

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
54 Stress and deprivation during growth and development and adulthood

reflects the deterioration of health in many of these settings (Cohen et al., 1994; Martin et al.,
1991; Milner & Smith, 1990; Pfeiffer & Fairgrieve, 1994; Stodder, 1994; Stodder & Martin,
1992).
While the record of enamel defects has been documented in populations undergoing the
transition from foraging to farming, other dietary shifts reveal a consistent set of results
linking a reduction in stress with improvement in diet. For example, in the shift from hunting
and gathering to cattle pastoralism in early Holocene North Africa, there is a decline in dental
defects in the primary dentition. This decrease reflects increased access to fat and protein in an
increasingly marginal setting of the Sahara Desert (Stojanowski & Carver, 2011). Comparison
of Jomon hunter-gatherers with Yayoi farmers in Japan also shows a decline in enamel
defects (Temple, 2010; Temple & Larsen, 2013). Temple and Larsen (2013) argue that the
introduction of a predictable and renewable food source at an advanced stage of development
resulted in a decline in physiological stress and improved health.
The impact of the adoption of agriculture on the stress experience in earlier human
populations is revealed in a number of settings where changes in health and nutrition are
documented by archaeological and osteological means. At Dickson Mounds, multiple indica-
tors show increasing levels of nutritional stress and infectious disease (see previously; and see
Chapter 3). Increase in prevalence of enamel defects – both macrodefects and microdefects – is
consistent with these findings (Goodman, 1989; Goodman & Armelagos, 1988; Goodman
et al., 1980; Goodman, Lallo et al., 1984; Lallo & Rose, 1979; Rose et al., 1978). The mean
frequency of hypoplastic defects increased from 0.9 to 1.2 to 1.6 per individual in the Late
Woodland, Mississippian Acculturated Late Woodland, and Middle Mississippian periods,
respectively. For the same three periods, the frequency of individuals affected increased from
45% to 60% to 80%. Most defects occurred in juveniles during ages two to four years for the
first two periods. Defects in the late prehistoric intensive agriculturalists (Middle Mississip-
pian) were earlier (<2 years) than in either of the two previous periods, indicating that stress
occurred in earlier childhood as nutritional quality worsened and disease intensified.
Comparisons of age-at-death for Dickson Mounds individuals with and without hypo-
plasias reveals that in the final two prehistoric periods, mean age-at-death of individuals
without hypoplasias is 36.6 years and 37.5 years, and mean age-at-death for individuals with
hypoplasias is 31.3 and 30.2 years (Goodman, Lallo et al., 1984). These results are similar to
other settings showing an inverse relationship between age and presence of enamel defects –
adults have fewer defects than juveniles, older juveniles have fewer defects than younger
juveniles, and/or older adults have fewer defects than younger adults (Boldsen, 2007; Cook,
1990; Cook & Buikstra, 1979; Duray, 1996a, 1996b; Irei et al., 2008; King et al., 2005;
Oxenham, 2006; Rose et al., 1978; Saunders & Keenleyside, 1999; Simpson et al., 1990;
Stodder, 1997; Swärdstedt, 1966; White, 1978). These findings are consistent with those in
living populations that individuals experiencing a greater number of stress episodes during
childhood are predisposed to early death, a finding consistent with the early developmental
origins hypothesis regarding the clear association between stress in early life (prenatal and
postnatal) and early mortality and compromised health in later life (and see Armelagos et al.,
2009; Beltrán-Sánchez et al., 2012; Temple, 2014).
Wilson bands in archaeological remains also provide an important avenue for investigating
other major transitions in human populations that might compromise health status and stress
levels. Wright (1990) determined prevalences of bands in precontact and contact-era

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.5 Skeletal and dental pathological markers of deprivation 55

mandibular canines from Lamanai, Belize, a Maya center occupied from the Preclassic
through Historic periods. Following an initial period of abandonment after contact by
Europeans, the site was reoccupied as a Catholic mission by Mayan Indians until the mid-
seventeenth century. Diet changed relatively little in the prehistoric to historic transition, but
archival records indicate that other stressors (e.g., European-introduced diseases) comprom-
ised health in native populations following contact. Comparisons of Postclassic and Historic
dentitions show a dramatic increase in physiological stress: 84% of bands observed in the
samples combined are from the Historic dentitions. Historic individuals also show more bands
than seen in precontact individuals (2.4 versus 0.88 per individual, respectively). Given the
lack of major dietary changes, Wright (1990) argues that the differences in microdefect
prevalence can be attributed to changing disease patterns with contact, such as the introduc-
tion of malaria and Old World parasitic infections, diseases leading to acute health crises.
Similarly, early colonial Native Americans in Spanish Florida show considerably greater
prevalence of Wilson bands than do their prehistoric predecessors (Simpson, 2001). Although
the elevation in stress can be attributed to a variety of environmental stressors associated with
Spanish missionization in this setting, a key factor was likely dehydration owing to diarrheal
disease and the reliance on a corn-based weaning diet.
A contrasting temporal trend in microdefect prevalence in comparison with prehistoric and
contact-era native populations is identified at another mission locality in Belize. Microdefect
prevalence in mission Indians at Tipu decreases in comparison with precontact Indians (Cohen
et al., 1994; Danforth, 1997). The different temporal trends at Lamanai and Tipu may indicate
very different contact experiences within the relatively small geographical setting of Belize.
Lamanai served as an important center where populations were relocated from nearby villages
to the town, whereas Tipu was only marginally affected by population relocation and concen-
tration during the mission period. Thus, the contrasting contact experiences at Lamanai and
Tipu may explain the different patterns of morbidity.
Temporal patterns documenting increase in stress are also displayed in foraging popula-
tions undergoing significant adaptive shifts. In the Santa Barbara Channel Islands region,
there is a marked increase in frequency of hypoplasias in the Middle period when populations
underwent a transition from hunting and gathering of terrestrial foods to a heavy reliance on
marine foods, especially fish (Lambert, 1993). Hypoplasias in the mandibular left canine
increased in prevalence from 18.4% in the Late Early period to 49.2% in the Early Middle
period. This increase in stress mirrors temporal changes observed for other nonspecific
indicators (e.g., periosteal reactions), all of which appear to be associated with problems
relating to increasing sedentism, population aggregation, and declining resource availability.
Enamel defect prevalences appear to differ between class or social groupings in a number of
settings, suggesting that higher-status individuals have better diets or other positive environ-
mental factors than lower-status individuals. Prehistoric Rudston and Burton Fleming (Eng-
land) and Medieval-period Westerhus (Sweden) and Koksijde (Belgium) populations represent
unambiguous examples of status differences in enamel defect prevalence, with higher-status
individuals showing considerably fewer enamel defects than lower-status individuals (Peck,
2013; Polet et al., 2000; Swärdstedt, 1966) (Figure 2.18). Thus, in these settings, during the
years of growth and development of the dentition, higher-status juveniles may have enjoyed
better health than did lower-status juveniles. Similarly, in the Dickson Mounds series and in
the Pete Klunk Mound group in the lower Illinois River valley, high-status individuals have

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
56 Stress and deprivation during growth and development and adulthood

High Status

Prevalence (%) of enamel hypoplasias


100
Intermediate Status

80 Low Status

60

40

20

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6 6.5 7 10-17
Age at tooth mineralization (years)
Figure 2.18 Prevalence (%) of enamel hypoplasias by social groups of high, intermediate,
and low status at Westerhus, Sweden. (Adapted from Swärdstedt, 1966.)

fewer hypoplasias than low-status individuals (Cook, 1981a; Goodman & Armelagos, 1988),
suggesting a possible association between status and stress history in both settings. In
Lambayeque, Peru, and Copán, Honduras, lower-status individuals have a greater prevalence
of enamel defects than higher-status individuals (Shimada et al., 2004; Storey, 2005). The
similar pattern of age differences in vertebral neural arch size, tooth size, and enamel defect
frequency, especially in the comparison of juveniles and adults from the same population,
strongly suggests that individuals surviving to adulthood enjoyed relatively better health than
did those members of the population who expired prior to reaching adulthood.
Enamel defects are not a health risk per se, but abundant clinical evidence indicates that
enamel defects – hypoplasia and hypocalcification – predispose teeth to cariogenesis
(Baume & Meyer, 1966; Mellanby, 1934; Nikiforuk & Fraser, 1984; and see review in Duray,
1992). Called “circular caries,” the association between enamel defects and caries has been
documented in Middle Woodland and Late Woodland deciduous teeth in lower Illinois River
valley (Cook & Buikstra, 1979). Both periods contain high prevalences of circular caries, but
there is an especially high prevalence in the later period when maize agriculture intensified
(Cook and Buikstra, 1979). Follow-up study of microdefects from this setting indicates a close
association between circular caries and microdefects (Cook, 1990). Identification of age-
of-occurrence of stress indicators reveals that physiological perturbations are perinatal,
reflecting poor health of both the mother and her infant around the time of birth.
Dentitions from the Libben site, Ohio, show a gradient in caries susceptibility in comparison
with different types of gross enamel defects – teeth with gray-chalky hypocalcifications are
more carious than discolored teeth (Duray, 1990, 1992). These associations indicate that
weakened enamel structure promotes cariogenesis. In contrast to many other prehistoric
foragers, the Libben series displays high caries prevalence. Duray (1990) suggests that high
caries prevalence can be attributed to the high prevalence of hypocalcifications. In contrast to

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.6 Adult stress 57

the association between caries and hypocalcification, teeth with linear hypoplasias appear to
be caries-resistant. Duray (1990) speculates that hypermineralization of the defect may
suppress cariogenesis.
Hypocalcifications and dental caries in deciduous teeth have also been identified in the
Classic-period Maya Indians from Copán, Honduras (Storey, 1992b, 1992c), and in native
populations from the Mariana Islands, Polynesia (Hanson, 1990). In both settings, native
twentieth-century populations have high prevalences of circular caries. In Mesoamerica, these
prevalences are linked with over-reliance on carbohydrates, poor nutrition, and the synergis-
tic relationship between diet and disease (Storey, 1992c). In the Mariana Islands, unusually
high levels of circular caries may be related to a number of local conditions, including
excessive fluoride intake by pregnant and lactating women, poor water quality, the consump-
tion of highly cariogenic starch diets with weaning, and specific nutrient deficiencies (e.g.,
protein) in mothers and their infants (Hanson, 1990).
Sex differences in prevalence of hypoplasias and other enamel defects in archaeological
series are highly variable. For example, no differences in hypoplasia prevalence are present
between adult females and males in the Dickson Mounds series (Goodman et al., 1980; and see
discussions of other North American samples by Danforth, 1997; Danforth et al., 1997;
Lanphear, 1990; Martin et al., 1991; Powell, 1988; Stodder, 1997; Wright, 1990).
Clinical, epidemiological, and bioarchaeological studies indicate that enamel defects are
more common in males than in females, more common in females than in males, and in equal
prevalences (Goodman & Rose, 1990; Goodman et al., 1991; King et al., 2005). At least one
review suggests that stress buffering, often associated with females, does not influence defect
expression to any appreciable degree (Guatelli-Steinberg & Lukacs, 1999). Immature females
have more hypoplasias than immature males in Tezonteopan, Mexico, which is consistent
with other evidence indicating worse nutrition in girls than in boys in this setting (Goodman
et al., 1991). Hypoplasia prevalence does not differ between South Asian males and females in
settings where daughter neglect results in greater malnutrition and mortality in females, at
least for some regions (Lukacs and Joshi, 1992). The differences in findings between Mexico
and South Asia suggest that differential treatment of males and females during childhood will
not necessarily be reflected in differences in enamel defect prevalences.

2.6 Adult stress


2.6.1 Bone mass
As during the juvenile years, cortical bone continues to be a highly dynamic tissue in
adulthood. Bone apposition takes place on the endosteal and subperiosteal surfaces well into
the third or fourth decades of life, peaking at about the age of 35 years (Garn, 1970; Heaney,
1993; Pfeiffer & Lazenby, 1994). Universally, by about 40 years of age, bone commences
resorption endosteally but continues to be deposited periosteally. The imbalance of bone loss
(most of which is endosteal) and bone gain (mostly periosteal) results in a net reduction of
bone tissue during and following the fifth decade of life (Bales & Anderson, 1995; Frost, 2003;
Garn et al., 1967, 1992; Grynpas, 2003; Smith & Walker, 1964) (Figure 2.19).
Systematic adult bone loss severe enough to result in increased fragility and risk of fracture
in older adulthood is a complex disorder called osteoporosis (Agarwal, 2008; Anderson, 1995;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
58 Stress and deprivation during growth and development and adulthood

Figure 2.19 Left to right: sequential loss of adult bone with age, in second metacarpals,
Post-Medieval St. Bride’s Lower Churchyard, London. (Brickley & Ives, 2008; reproduced
with permission of the Museum of London.)

Anderson & Pollitzer, 1994; Glencross & Agarwal, 2011; Heaney, 1993; Stini, 1990, 1995).
Depending on factors relating to muscle mass, activity level, pregnancy, lactation, and diet,
bone loss after the age of 40 years is substantial, amounting to 1% – 5% per year. The
identification of variation in populations of disparate ancestry also points to genetic factors
that predispose some groups to osteoporosis (Anderson & Pollitzer, 1994; Cho et al., 2006;
Nelson et al., 2004). Two types of bone loss due to osteoporosis are identified clinically,
including that arising from reduction in estrogen levels following menopause (Type I), and
gradual age-related reduction in bone mass in adult females and males (Type II) (Drezner,
1995; Stini, 1990). Women lose relatively more bone mass than do men, due to the combined
effects of Type I and Type II osteoporosis. Estrogen is critical in bone maintenance (Drink-
water, 1994). The reduction in estrogen can even result in significant bone loss for younger
women undergoing overly vigorous exercise regimens and accompanying loss of menstru-
ation (secondary amenorrhea) (Anderson, 1995; Kreiner, 1995).
The rate of bone loss in human populations is variable, and environmental factors such as
nutritional status are significant influences (Arnaud & Sanchez, 1990; Martin et al., 1985;
Nelson et al., 2004; Pollitzer & Anderson, 1989; Schaafsma et al., 1987). Clinical evidence

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.6 Adult stress 59

indicates that individuals with low calcium intakes are more prone to adult bone loss, and
other dietary factors such as high protein consumption are also implicated (Arnaud &
Sanchez, 1990; Nordin et al., 1993; Pfeiffer & Lazenby, 1994; Stini, 1990, 1995). Body weight,
heredity, and lactation status are also important risk factors (Agarwal & Stout, 2003; Arnaud &
Sanchez, 1990; Evers et al., 1985; Heaney, 1993; Kreiner, 1995; Pollitzer & Anderson, 1989;
Schaafsma et al., 1987; Stini, 1990, 1995). Comparisons of active versus sedentary popula-
tions or athletes versus non-athletes indicate the strong influence of physical activity on bone
maintenance: simply, active individuals have stronger, denser bone than sedentary individ-
uals (Anderson & Pollitzer, 1994; Drinkwater, 1994; Kohrt et al., 2004; Lacey et al., 1991;
Marcus et al., 1992; McMurray, 1995; Yano et al., 1984). Due to a decrease in physically
demanding lifestyles in contemporary settings (e.g., Sweden, United States, United Kingdom,
China), there appears to be a rapid secular increase in osteoporosis in general and osteoporotic
fractures in particular, matching the remarkably rapid changes in lifestyle (Allander, 1995;
Kohrt et al., 2004; Lau et al., 1990; Ruff, 2006).
Adult bone mass is documented in human remains from a variety of archaeological settings.
Much of this research shows either a general similarity or accelerated patterns of bone loss in
archaeological samples and in living populations (Carlson et al., 1976; Cho & Stout, 2003; Cook,
1984; Dewey et al., 1969; Robling & Stout, 2003; Van Gerven, 1973). Variation in relation to
differing lifestyles and subsistence strategies has been examined in some detail, using alterna-
tive data collection protocols, including raw measures of bone mass (cortical thickness [CT],
cortical area [CA], bone mineral content [BMC]), or size-standardized measures (percentage of
cortical area [%CA or PCCA] or percentage of cortical thickness [%CT or PCCT]) (Ruff, 1992).
Comparisons of femoral cortical thickness in X-group (AD 350 – 550) intensive agriculturalists
from the Wadi Halfa area of Sudanese Nubia with modern Euroamericans and Native Americans
reveal similar trends of initial gains in bone mass from the third to fourth decades, followed by
losses (Martin & Armelagos, 1979; Martin et al., 1985). In Nubian females, bone mass decreased
after 20 years of age. Martin and coworkers (1985) speculate that premature osteoporosis was
due to nutritional inadequacies associated with an over-reliance on a single dominant crop
(millet), such as protein – calorie malnutrition or imbalance of calcium – phosphorus ratios, or
the influence of disease. Perhaps the bone losses in childhood in this setting (Hummert, 1983;
and see earlier) predisposed adults, especially females, to premature bone loss.
Similarly, bone mass (percentage of cortical area from second metacarpals) in late prehis-
toric maize agriculturalists from southern Ontario is below what has been documented in
living populations (Pfeiffer and King, 1983; and compare with Garn, 1970). Although various
factors may be involved, the reliance on maize and attendant protein – calorie malnutrition
may have contributed to low bone mass.
Based on the assumption that cortical thickness and Nordin’s Index (cortical thickness/total
subperiosteal area) are useful measures of bone mass and nutritional quality, Owsley (1991)
compared femoral bone mass in a temporal series of Great Plains Arikara dating from c. AD
1600 – 1832. These comparisons reveal an increase in bone mass in the transition from the
late prehistoric to early protohistoric period in the late seventeenth century, which Owsley
regards as “a positive change in nutritional status,” perhaps relating to increased availability
of protein acquired through hunting and trade (1991:109). By 1800, bone mass declined
dramatically. Owsley suggests that these declines are due to the stresses associated with

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
60 Stress and deprivation during growth and development and adulthood

biological and social disruptions of disease, warfare, and other negative environmental
circumstances in the early nineteenth century.
In contrast to the Arikara, no temporal change in bone mass (Nordin’s Index) could be
detected in a sequence of human remains from the lower Illinois River valley dating from the
Archaic to Mississippian period (Cook, 1984). This suggests that the profound change in diet
involving the shift to maize agriculture in later prehistory had no bearing on bone mainten-
ance. Some diseased individuals in the Mississippian period expressed relatively low bone
mass. Individuals with skeletal tuberculosis had remarkably low bone mass, suggesting that
non-dietary factors play an important role in bone maintenance in this region.
Most bioarchaeological research emphasizes the direct role of diet and nutrition in explain-
ing variation in bone maintenance. This perspective ignores the important influence of
mechanical loading and activity on bone mass and how it is distributed (and see Chapter 6).
Continued subperiosteal expansion in adults compensates for medullary expansion and
endosteal bone loss in order to maintain the mechanical integrity of the bone cross-section
under loading regimes. The raw measures of bone mass frequently used in studies of archaeo-
logical remains may present an incomplete picture of bone remodeling and health status (Ruff,
2008; Ruff & Larsen, 1990). Thus, low bone mass does not necessarily indicate inadequate
bone mass (Pfeiffer & Lazenby, 1994).

2.6.2 Skeletal histomorphometry


Primary among microscopic structures in remodeled human bone are multicellular features
called secondary osteons (primary osteons are found in primary, unremodeled bone; see
Martin et al., 1998). Secondary osteons are created in a two-stage remodeling process: first,
osteoclasts destroy existing bone tissue, resulting in minute tunnels or resorption spaces; and
second, unmineralized bone (osteoid) is deposited by osteoblasts and subsequently mineral-
ized in a series of incremental layers on the surfaces of these tunnels. The fully formed osteon
consists of a series of concentric bone layers organized around vascular canals called
Haversian canals (Crowder & Stout, 2011; Martin et al., 1998; McLean & Urist, 1968)
(Figure 2.20). In normal, healthy individuals, remodeling is a uniform process that begins in
early childhood and continues throughout life. Remodeling rates are influenced to a large
degree by a variety of diseases and nutritional disorders (Frost, 1966, 2003), but activity is also
an important mediating factor (Burr & Allen, 2013; Cho & Stout, 2003). These rates can be
quantified in recent and archaeological remains, thus representing an indicator of stress and
lifestyle history as it relates to health of bone tissue (Agarwal, 2008; Cho & Stout, 2003;
various in Crowder & Stout, 2011; Martin & Armelagos, 1985; Mulhern, 2000; Mulhern & Van
Gerven, 1997; Simmons, 1985; Stout & Simmons, 1979). For example, various measures of
bone remodeling rates for ribs collectively show less bone remodeling per unit area in the
Roman-era Isola Sacra, Italy population than in modern European Americans, indicating
generally healthier bone and thus less risk of fracture and fragility in the former than in the
latter (Cho & Stout, 2003) (Figure 2.21).
Comparisons of prehistoric populations from the lower Illinois River valley (Gibson, Ray,
and Ledders sites), Florida (Windover site), and Peru (Paloma site) show that maize agricul-
turalists (Ledders) have greater remodeling rates than hunter-gatherers (Stout, 1978, 1983;
Stout & Lueck, 1995; Stout & Teitelbaum, 1976). The greater remodeling rates in the maize

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.6 Adult stress 61

(a)

Periosteum

Haversian systems
Lamella

Vein
Lacuna
Lymph vessel

Artery

(b)

Figure 2.20 Schematized (a) and actual (b) cross-section of compact bone showing major
microstructures discussed in the text. The ring-like structures (Haversian systems or
osteons) house the Haversian canals. Each lacuna contains cells that maintain the bone
tissue and are arranged in layers (lamellae). ([a] Illustration by Dennis O’Brien; Larsen,
1987. © Academic Press, Inc.; [b] photograph by Samuel Stout.) (A black and white version
of this figure will appear in some formats. For the color version, please refer to the plate
section.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
62 Stress and deprivation during growth and development and adulthood

80

70 Isola Sacra

60 European American

50

40

30

20

10

0
Tt.Ar Ct.Ar En.Ar OPD
(mm2) (mm2) (mm2) (#/mm2)

1.4

1.2

0.8

0.6

0.4

0.2

0
r x 2 ) r) r) 2 )
t.A de m 2 /y 2 /y m
/T in m m m /m
Ar ic r( m /m
2
C
t. bol n.
A (#
/ 2 m
ra f m (m
O c. (m
Pa A R BF
R
BF et
N
Figure 2.21 Comparison of mean histomorphometric variables: Tt.Ar, total cross-sectional
area; Ct.Ar, cortical bone area; En.Ar, endosteal area (bone marrow cavity area); OPD,
osteon population density; Ct.Ar/Tt.Ar, relative cortical area; Parabolic index, optimal
distribution of cortical bone around marrow space; On.Ar, average cross-sectional area
with minimally 25 complete osteons per section; Ac.f, mean activation frequency
associated with the number of osteons annually created per area unit; BFR, average bone
formation rate; Net BFR, new bone formation rate. (Adapted from Cho & Stout, 2003;
reproduced with permission of authors and Kluwer Academic/Plenum Publishers.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.6 Adult stress 63

agriculturalists may reflect the effects of nutritional stress associated with maize-based, low-
protein diets (Pfeiffer & Lazenby, 1994; Stout, 1983), or perhaps, variation in skeletal tissue
maturation rates in different populations (Stout & Lueck, 1995).
The rate of mineralization of osteoid (unmineralized bone matrix) in the Haversian canal
can be influenced by systemic stress. Under conditions of normal bone development, osteons
mineralize uniformly, but under conditions involving slower growth – such as with nutri-
tional stress or disease – delayed osteoid mineralization results in the creation of hyperminer-
alization zones. Viewed in cross-section, one or more hypermineralized zones appear as rings
of increased density comparable to lines of growth disruption found at the ends of long bones
(Bartsiokas & Day, 1993; Martin & Armelagos, 1985; Simmons, 1985; Stout & Simmons,
1979). The presence of these “double-zone” osteons indicates the presence of physiological
stress and growth disruption. The frequency of double-zone osteons per unit area of bone is
positively correlated with the total amount (cortical area) of bone in the Nubian skeletons
(Martin & Armelagos, 1985). This may indicate that some critical threshold of metabolic
activity is essential for recovery from growth arrest (Martin & Armelagos, 1985).
Histomorphometric study of three populations with varying subsistence strategies –
primarily meat (Alaskan Eskimo), mixed foraging and farming (Arikara), and intensive maize
agriculture (Pueblo) – shows a high degree of variability in osteon structure. Osteons in a
number of individuals in these samples contain second and smaller remodeling sequences
(Eriksen, 1980; Richman et al., 1979). Called Type II osteons, these structures represent sites of
accelerated availability of calcium. Alaskan Eskimos have the highest frequency of Type II
osteons, perhaps reflecting heavy reliance on meat in comparison with more plant-oriented
Arikara and Pueblo Native Americans.
The study of histomorphometrics in adult Nubian remains provides important comple-
mentary information to the aforementioned analyses of bone mass. Nubian adult femora
possess abnormally large, active resorption spaces, which may result from nutritional stress
due to agricultural dependence (Martin and Armelagos, 1979, 1985). The overall quality of
bone is reduced in these groups due to the high degree of inadequately mineralized, porous
cortex. These findings underscore the general observation that increased porosity is due, at
least in part, to histological changes, including enlarged resorption spaces. Moreover, faster
remodeling will yield more osteons and thus more Haversian canals, contributing to increased
intracortical porosity. The role of mechanical environment is suggested strongly by variation
in microstructure, however. In this regard, adult females have larger femoral osteons than
adult males (Mulhern and Van Gerven, 1997), which reflects less mechanical demand on
women than men. This is consistent with the observation that osteon size is not different in
ribs analyzed in adult males and females because ribs are non-load bearing (Mulhern, 2000).

2.6.3 Vertebral trabecular architecture


Issues of bone loss and remodeling have been documented in the archaeological record via the
study of trabecular architecture in adult vertebrae. Comparison of individuals from the
Medieval Wharram Percy population in Britain reveals a clear trend of age-related trabecular
bone structure associated with bone remodeling and loss, occurring primarily before middle
age (Agarwal et al., 2003, 2004). Unlike modern populations that generally show continued
trabecular structure change, the Medieval sample showed stasis in postmenopausal life.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
64 Stress and deprivation during growth and development and adulthood

Although the reasons for this difference with modern populations are not known, the greater
mechanical demands and differences in lifestyle in general may explain the maintenance of
structural integrity in older adults at Wharram Percy in contrast to living societies where
adults are less active (Agarwal et al., 2004).

2.7 Summary and conclusions


The sensitivity of the human skeleton to impoverished environments, especially during the
years of growth and development, is revealed by the study of a range of stress indicators,
including growth rates, attained adult stature, pelvic inlet shape, vertebral neural canal size
and shape, tooth size, tooth size asymmetry, and various skeletal and dental pathological
conditions. Once adulthood is reached, fewer changes arising from physiological stress are
exhibited in the hard tissues. Bone loss and growth arrest in developing osteons as well as
abnormal remodeling rates are highly informative about stress history. Most skeletal and
dental stress indicators reflect episodes of physiological perturbation during childhood;
nevertheless, their measurement and interpretation serve as an indication of the stress experi-
ence for the population generally. Studies of stress based on archaeological bones and teeth
reveal a number of consistent patterns. Under circumstances conducive to increased stress –
such as poor nutrition, population aggregation, and increased infectious disease – skeletons
and dentitions exhibit stress indicators in elevated prevalences.
In a variety of disadvantaged circumstances, juveniles exhibit growth reduction. This
reduction is commonly tied to chronic nutritional deficiencies often resulting from the
synergy between poor nutrition and infection. If these conditions are maintained throughout
the years of growth and development, the affected individuals are likely to be short-statured
as adults. Overall, the early environment has long-lasting effects on health status over the
lifespan. Juveniles experiencing elevated stress can have poor health and shortened lifespans
as adults (and see Henry & Ulijaszek, 1996). Simply, elevated stress early in life due to various
environmental factors – poor nutrition, disease, and socioeconomic stressors – contributes to
early death and reduced longevity.
Thus, the composite of stress described in this chapter shows a tendency for increased
indicator prevalence in contexts where environmental, social, and other perspectives are
creating a poor living environment. It is important to point out, however, that within specific
settings, there is not a universal design of simultaneous patterns of increasing stress as based
on skeletal and dental evidence. For example, in early and late Medieval populations from
England, the comparison of individuals having hard tissue evidence of elevated stress,
minimal stress, and no stress, based on relative prevalence of hypoplasias and cranial
porosities (cribra orbitalia, porotic hyperostosis), showed no difference in long bone lengths
in the three groups (Ribot & Roberts, 1996; and compare with Pinhasi et al., 2014; Temple,
2008). In contrast, in the American Southwest, there is an association between growth and
cranial porosities: young children (birth to five years old) with cranial porosities are growth
stunted (Schillaci et al., 2011).
Do small-bodied humans have an adaptive advantage over large-bodied humans living
under circumstances of reduced resource availability or inadequate nutrition? Seckler (1980,
1982) proposed that shortness in body height in developing nations is an adaptation to
reduced food supplies. This reduction, he argues, results in individuals who are small, but

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
2.7 Summary and conclusions 65

healthy. If reduced body size is adaptive, then reduced height should have no associated
functional or demographic costs. In fact, a significant body of literature from multiple
disciplines illustrates that small body size is linked with various negative factors, including
increased disease and poor nutrition. Poor growth status is associated with a range of
functional costs and consequences, including decreased activity and poorer learning (Crooks,
1995; Dasgupta, 1993; Goodman, 1991, 1994; Stinson, 1992, 2000). Although smaller body
size appears to enable individuals to perform some activities with lower energy requirements,
the efficiency is reduced (Stinson, 1992). Clearly, there are negative consequences of small
body size in disadvantaged settings, indicating that this reduction is associated with elevated
morbidity, and is in fact maladaptive.
In the study of past societies, it is possible to test the “adaptive” versus “stress” hypotheses
via observation and quantification of the growth and nutrition stress indicators discussed in
this chapter. Significantly, various indicators are linked with decreased survival as determined
by mean age-at-death of individuals with and without (or with relative differences in
prevalence of ) the indicator. Enamel defects (macro- and microdefects), vertebral neural canal
size, height, and tooth size show clear links with lifespan. Where age-at-death and stress
indicators are examined concurrently, individuals without enamel defects or anemia-related
lesions, with larger neural canal size, with greater height, and with larger tooth size died later
in life. These findings suggest that skeletal stress indicators are related to quality of life and do
not represent adaptations.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:31:54, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.004
3 Exposure to infectious pathogens

3.1 Introduction
For the entire evolution of our species, we have been exposed to a wide range of infectious
agents – parasites, bacteria, and viruses – resulting in a range of diseases. The dental and
skeletal evidence for some of these diseases, mostly chronic, is well documented (Aufderheide
& Rodríguez-Martín, 1998; Ortner, 2003; Roberts & Manchester, 2005). Current bioarchaeo-
logical inquiry emphasizes biocultural perspectives of disease in relation to social, cultural,
and environmental contexts and risks for infection, impacts on population, and implications
for pathogen – host evolution (Armelagos & Van Gerven, 2003; Buikstra, 2010; Buikstra &
Cook, 1980; Larsen & Walker, 2010). This growing record of health in past populations and
especially the emphasis on origins and evolution of infectious disease in the biocultural
context provides a powerful approach for understanding health determinants and outcomes
in the world we live in today.
Infection by a pathogen does not always result in disease. The progression from infection to
disease depends on agent pathogenicity, transmission route from agent to host, and the
strength and nature of the response of the host (Brown et al., 2011; Inhorn & Brown, 1990;
Smith & Moss, 1994). Many acute infectious diseases or epidemics result in death of the
infected individual soon after microbial attack. These infectious diseases leave no skeletal
record, clouding the full picture of disease and its relationship to mortality in past popula-
tions. A number of chronic infectious diseases affect osseous tissues in patterned ways.
Despite the interpretive drawbacks and sometimes lack of specificity, the study of bone lesions
documenting disease provides important perspectives on health in earlier societies and the
impacts of particular living circumstances on the human condition.
The frequency of members of a population affected by a disease forms the baseline of
information from which to interpret health status and factors that influence it. Various means
of data presentation are used, but incidence and prevalence are most commonly reported in
clinical and epidemiological studies (Keyserling, 1988; Waldron, 1994). Incidence is defined as
the number of new cases in a population in a given period. Because the number of new cases
cannot be identified in archaeological settings with certainty (if at all), it is not possible report
on the incidence of a particular disease or pathological conditions. Prevalence, however, can be
observed in archaeological contexts because it represents the proportion of the population
affected by the disease at a single moment or within a period. Prevalence, therefore, provides an
avenue for addressing how common pathological conditions relating to disease occur under a
given set of circumstances, including geographical and temporal variation, sex, age, status and
socioeconomic position, residence, and other key contexts. This paleoepidemiological approach
is fundamental to understanding life conditions in past populations (Boldsen & Milner, 2012).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.2 Dental caries 67

This chapter focuses on disease prevalence as expressed in teeth and bones. Ortner (2003)
provides details on a wide variety of infectious diseases identified in archaeological remains
from around the world. This discussion of infection pertains to dental caries, periodontal
disease, antemortem tooth loss, nonspecific periostitis, and specific infectious
diseases – treponematosis, tuberculosis, leprosy, and vectored infections, especially bubonic
plague. These conditions are among the most frequently studied, they have distinctive
pathogenesis involving bone (except for bubonic plague, which leaves no skeletal signature),
and the diagnostic tools (e.g., differential diagnosis of skeletal lesions, ancient DNA docu-
mentation of pathogens) are providing new perspectives on the occurrence and impact of
infectious disease on humankind.

3.2 Dental caries


3.2.1 Description and etiology
Contrary to that which is often presented by anthropologists, the term “dental caries” does not
refer to lesions in teeth resulting from invasion of microorganisms. Rather, dental caries is a
disease process characterized by the focal demineralization of dental hard tissues by organic
acids produced by bacterial fermentation of dietary carbohydrates, especially sugars and (to a
lesser extent) starches. Dental caries is manifested in various states, ranging from slight enamel
opacities to extensive cavitation involving partial or complete destruction of tooth crowns and
roots (Hillson, 2008) (Figure 3.1). Because carious lesions (“cavities”) are readily observable in
both archaeological series and living populations, there is an abundance of published data on
prevalence for a variety of temporal and geographic settings around the globe.

Figure 3.1 Mandibular carious lesions; Ochsenfurt, Germany. Root caries have affected the
cement–enamel junction of the molar (right) and resulted in the complete destruction of the
tooth crown (left). (Photograph and copyright by Leslie L. Williams; reproduced with
permission.). (A black and white version of this figure will appear in some formats. For the
color version, please refer to the plate section.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
68 Exposure to infectious pathogens

Several essential and modifying factors are involved in the pathogenesis of dental caries.
The essential factors include: (1) teeth the surfaces of which are exposed to the oral environ-
ment; (2) the presence of aggregates of complex indigenous oral bacterial flora (especially
including but not limited to Streptococcus mutans, Lactobacillus acidophilus), salivary glyco-
proteins, and inorganic salts adhering to the tooth surfaces (called dental plaque); and (3) diet
(Burne, 1998; Hara & Zero, 2010; Rowe, 1982; Tanner et al., 2012). Modifying factors are
those that influence the site distribution and rate of carious lesion development. Such factors
include, but are not limited to, crown size and morphology, enamel defects, occlusal surface
attrition, tooth structure, food texture, oral and plaque pH, speed of food consumption, some
systemic diseases, age, child abuse, heredity, salivary composition and flow, nutrition, perio-
dontal disease, enamel elemental composition, and presence of fluoride in drinking water, and
other geochemical factors (Bowen, 1994; Burt & Ismail, 1986; Burt & Pai, 2001; Geddes, 1994;
Greene et al., 1994; Hara & Zero, 2010; Hildebolt et al., 1988, 1989; Hillson, 2008; Hunt et al.,
1992; Leverett, 1982; Maat & Van der Velde, 1987; Meiklejohn et al., 1992; Milner, 1984;
Molnar & Hildebolt, 1987; Molnar & Molnar, 1985b; Newbrun, 1982; Powell, 1985; Rowe,
1982; Woodward & Walker, 1994). Preliminary analysis of genes involved in taste, enamel
formation, and saliva composition suggests a genetic component associated with increased
risk of contracting dental caries (Wendell et al., 2010; Wright, 2010).
Intrinsic characteristics of food, its consistency, and the manner in which it is prepared
influence cariogenesis in human populations. A range of human and animal investigations
reveal that carbohydrates – sugars and starches – are central elements promoting dental caries
(Hillson, 2008). The relative amount of carbohydrates consumed and reliance on domesticated
plants largely explains the variation in prevalence of dental caries in archaeological and other
settings. As succinctly stated by Hara and Zero (2010:459), diet is “the main driver of the
caries process.”
The degree and rate of occlusal surface wear appears to be a mitigating factor in caries rates
in archaeological settings. Maat and Van der Velde (1987) found a negative correlation
between frequency of occlusal surface caries and degree of dental wear in molars from sailors
recovered from a seventeenth- and eighteenth-century Dutch whaling station in the Spitz-
bergen Archipelago (Svalbard). In this series, increased wear appears to be associated with
fewer carious lesions. They concluded that “(t)hese findings strongly suggest a competitive
relationship between progress in caries and attrition” (Maat & Van der Velde, 1987:281). The
studies linking reduced wear to increased caries prevalence are convincing, especially because
common sites of plaque and cariogenesis are the grooves and fissures of unworn crowns (and
see Christopherson & Pedersen, 1939; Corbett & Moore, 1976; Hillson, 2008; Milner, 1984),
especially of the posterior teeth, which have a differential (higher) risk of developing carious
lesions than the anterior teeth.
The relationship between low caries rates and high occlusal-wear environments should not
be overly generalized, however. In their study of the Mesolithic dentitions from Cabeço da
Arruda and Moita do Sebastiao, Portugal, Meiklejohn and coworkers (1992) found a positive
correlation between caries and wear in molars; the most heavily worn crowns are the most
carious. In this setting, the Mesolithic Portuguese individuals consumed figs and carob, foods
high in sugar content that also produce high rates of wear. Occlusal surface wear is also
excessive in Archaic-period foragers from the lower Pecos region of southwestern Texas,
resulting in pulp cavity exposure and tooth loss (Hartnady & Rose, 1991; Sobolik, 1994).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.2 Dental caries 69

Caries prevalence is high (14% of teeth) and is indistinguishable from many agricultural
groups (Larsen et al., 1991).
Coprolite analysis reveals that various highly abrasive materials were included in foods
consumed, including phytoliths, seeds and small bones, and calcium-oxylate crystals from
succulents and cacti (Hartnady & Rose, 1991; Sobolik, 1994). Historic accounts also document
the introduction of abrasives to food, including ash for sotal baking and dirt to “sweeten”
meals. High-carbohydrate foods such as succulent fibers, prickly pear fruits, pecans, and
mesquite resulted in active cariogenesis. Thus, like the Portuguese Mesolithic foragers, Pecos-
region Native Americans show a positive relationship between tooth wear and caries.

3.2.2 Temporal trends: foragers, farmers, and industrialized populations


The study of temporal trends of dental caries in archaeological samples has a long history.
Mummery (1870) was among the first to systematically document these trends in past
populations, observing increased caries prevalence in British populations. He related the
change in caries prevalence to cognitive development of children in a comparison of earlier
simple and later complex societies: “May we not therefore reasonably suppose that through
the diminished vitality consequent upon this diversion of the formative energy from the teeth,
by premature mental exertion, these organs necessarily become degenerated; and that this
circumstance constitutes one great difference between the teeth of the intellectual and those
of the uncultivated families of man” (1870:73).
It has become abundantly clear since Mummery completed his ambitious study that diet
and subsistence technology are far more important in understanding temporal variation in
caries prevalence. For many areas of the globe, there is a trend toward an increase in caries
prevalence or severity, or both, with intensified consumption of carbohydrates and decreased
consumption of protein (Burt, 1993; Mayhall, 1970; Moorrees, 1957; Oranje et al., 1935 – 37;
Pedersen, 1947; Pezo Lanfranco and Eggers, 2010; Price, 1936; Russell et al., 1961; Temple &
Larsen, 2007; Walker et al., 1998; and others) in past human groups.
Numerous investigators have detailed increases in carious lesion frequencies in prehistoric
agriculturalists compared to hunter-gatherers that preceded them. From a sample of popula-
tions drawn globally, Turner (1979) determined average frequencies of adult teeth affected
(incisors, canines, premolars, molars combined). His investigation revealed a steadily increas-
ing gradient for groups practicing foraging to an agricultural way of life: foraging, 1.7%;
mixed foraging/agriculture, 4.4%; agriculture, 8.6% (percentages calculated from Turner,
1979). A subsequent and even more in-depth study of caries frequency in a wide range of
populations globally showed wide variation in observed caries prevalence within and between
subsistence economies (Hillson, 2008; Lukacs, 2008). Nevertheless, the basic gradient from
low to high caries prevalence based on level of commitment to agriculture is clear, but the
degree of this rise varies, based on the level of commitment to particular agricultural products
(e.g., maize vs. wheat vs. rice).
Among some of the most comprehensive data on dental caries, prevalences are from the
Eastern Woodlands of North America. Figure 3.2 shows prevalence values, summarized
temporally, based on a sample of 180 archaeological populations. The dichotomy between
foragers and maize farmers is straightforward, but with overlap in values. The three forager
periods (Archaic to Middle Woodland) have less than 7% of carious teeth, and the three

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
70 Exposure to infectious pathogens

25
Eastern North America
21.0
Dental Caries
Carious Teeth (Percentage) 20
18.0

14.4
15

10

5.1
5 3.3
2.1

0
Archaic

Contact
Late Woodland
Early Woodland

Late Prehistoric
Middle Woodland
1000 BC 500 BC AD 500 AD 1000 AD 1500

Figure 3.2 Percentage of teeth affected by dental caries in eastern North America.
(Based on data from Larsen et al., 1991; Milner, 1984.)

agriculturalist periods (Late Woodland to Contact) have more than 7% of carious teeth. Within
the agriculturalist periods, there is a high degree of variability in caries prevalence. This
pattern mirrors observations of living populations with broadly similar diets (Hillson, 2008;
Walker et al., 1998). Relatively small differences in diet and food processing technology can
result in large differences in caries prevalence. Overall, there is a clear tendency for prehistoric
and early contact-era maize farmers to have higher caries prevalence than prehistoric foragers
(and see Gold, 2004; Lambert, 2000a; Larsen, 2014; Larsen et al., 1991; Milner, 1984; various
in Cohen & Armelagos, 1984; Cohen & Crane-Kramer, 2007; Steckel & Rose, 2002). The late
mission-period series from Amelia Island, Florida, shows an unusually high caries prevalence
for North America (Larsen et al., 1991, 2007). This is not unexpected given that missionization
was accompanied by considerable intensification in production and consumption of maize
during the sixteenth and seventeenth centuries in Spanish Florida (Hann, 1988; Jones, 1978;
Worth, 2001) and in other Spanish colonial settings (Bruwelheide et al., 2010; Klaus & Tam,
2010; Ubelaker, 1994).
The reasons for caries increase in Eastern Woodlands agriculturalists are varied and
complex, but the chief factor is related to the presence of sucrose in maize (Hardinge et al.,
1965). Because sucrose is a simple sugar, it is readily metabolized by oral bacteria. Declining
tooth wear in many settings was likely an important factor (see Chapter 7 for a discussion of
tooth wear change), but the common occurrence of maize consumption largely explains
these changes over this expansive area of North America. As such, the recognized
and dramatic positive association between maize consumption – in both its adoption and

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.2 Dental caries 71

intensification – and dental caries in eastern North America is useful for tracking temporal
changes in dental health in specific settings, where social, cultural, and dietary shifts are
documented using other independent lines of archaeological evidence (Cassidy, 1984; Cucina
et al., 2011; Danforth et al., 2007; Driscoll & Weaver, 2000; Gold, 2000; Hoyme & Bass, 1962;
Larsen et al., 1991, 2007; Milner, 1984; Patterson, 1984; Perzigian et al., 1984; Sciulli, 1997;
Sciulli & Oberly, 2002; Smith, 1982; Sullivan, 1990; and others). In at least one setting where
intensive farming was replaced by mixed foraging and farming, caries prevalence shows a
decline (Lambert, 2000b). Moreover, the range of variation is strongly influenced by geog-
raphy in this and other settings. For example, in the coastal plains of North Carolina and
Georgia, variation in caries prevalence reflects differing levels of commitment to maize
farming (Driscoll & Weaver, 2000; Hutchinson, 2002; Lambert, 2000b; Larsen et al., 2007),
a pattern consistent with dietary reconstruction based on carbon stable isotope ratios
(Hutchinson, 2002; Hutchinson et al., 1998; Larsen, Griffin et al., 2001; and see Chapter 8).
For many native populations of Eastern North America, maize agriculture was the first
significant experience with plant domestication. In several regions, it was not a new experi-
ence. In the North American midcontinental region (modern states of Ohio, Kentucky, and
Illinois), at least five native starchy plants were domesticated 2000 – 4000 years ago (Gremil-
lion, 2002; Smith & Yarnell, 2009). For this period, Rose and coworkers (1991) identified an
increase in carious lesion frequency that is likely linked with this reorientation of diet well
before the adoption of maize in later prehistory (beginning c AD 800 – 900).
There is limited understanding of the relative differences of the impact of agriculture on
oral health of populations consuming various types of plant domesticates. European and
western Asian Neolithic populations have generally lower caries prevalence values than those
of prehistoric maize farmers in North America (Eshed et al., 2006; Lubell et al., 1994). This
suggests that Neolithic domesticates were either less important or less cariogenic than maize
in New World populations. Moreover, in the Levant, there is no significant difference in the
pre-farming Natufian period (6.4% carious teeth) compared to the early farming Neolithic
period (6.7%) (Eshed et al., 2006). However, early and late Natufians have 1.5% and 6.8%
carious teeth, respectively (Eshed et al., 2006), suggesting that a greater reliance on plant
domesticates began before the Neolithic period.
By the same token, some Asian populations that consumed rice have relatively low caries
prevalence compared to maize-based populations, owing to apparently lower cariogenic
properties of rice (Sheiham, 2001; and see Domett, 2001; Domett & Tayles, 2007; Oxenham,
2006; Oxenham et al., 2006; Pietrusewsky & Ikehara-Quebral, 2006; Sakashita et al., 1997;
Tayles et al., 2000; Willis & Oxenham, 2013; but see Inoue et al., 1986; Temple & Larsen,
2007, 2013). In Southeast Asia, while there are temporal increases in rice production and
consumption, there are varying associated patterns of dental caries prevalence, including
decline (Domett, 2001), very little change (Pietrusewsky & Douglas, 2002a, 2002b), and
increase (Newton et al., 2013; Oxenham et al., 2006). These findings suggest that there is a
mixed pattern of health changes with intensification of rice agriculture (and see Clark et al.,
2014).
As a general trend, however, comparisons of pre- and postagricultural populations in a
wide range of settings reveal trends involving increases in caries prevalence irrespective of the
type of cultigens consumed, including East and South Asia (Douglas & Pietrusewsky, 2007;
Fujita, 1995; Inoue et al., 1986; Lukacs, 1990, 1992; Lukacs & Minderman, 1992; Lukacs et al.,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
72 Exposure to infectious pathogens

1989; Okazaki et al., 2014; Pechenkina et al., 2013; Temple & Larsen, 2007, 2013), the Middle
East (Littleton & Frohlich, 1993; Smith, Bar-Yosef et al., 1984; Smith & Horwitz, 2007), Europe
(Bennike, 1985; Bennike & Alexandersen, 2007; Brabant, 1967; Brinch & Møller-Christensen,
1949; Brothwell, 1959; Corbett & Moore, 1976; Cunha et al., 2007; Hardwick, 1960; Meikle-
john et al., 1984; Moore & Corbett, 1971, 1973, 1975; O’Sullivan et al., 1993; Roberts & Cox,
2003; Tóth, 1970; Wells, 1975; Whittaker, 1993), northeast Africa (Armelagos, 1969; Rose
et al., 1993; Smith & Jones, 1910), and South America (Alfonso et al., 2007; Pechenkina,
Vrandenburg et al., 2007; Ubelaker, 1984, 1994; Watson et al., 2010). In the rare instances
where populations shifted back to hunting and gathering, caries prevalence declined (Lukacs,
2007a).
Measurement of carious lesion size and location on tooth crowns provides an important
means of assessing the severity of the disease process. Temporal comparisons of foragers and
farmers, or earlier and later farmers, from a number of settings indicate an increase in lesion
size and shift in the distribution on tooth crowns from mostly the occlusal surface to other
surfaces (e.g., roots) (Larsen, 1982; Pezo Lanfranco & Eggers, 2010; Smith, 1982).
Elevated prevalences or severity of dental caries are not limited to groups relying on
domesticated plants. Prehistoric foragers from the lower Pecos River valley and central Texas
have values that are well within the ranges reported for agriculturalists (Bement, 1994;
Hartnady & Rose, 1991; Marks et al., 1988; Sobolik, 1994). In these settings, the consumption
of sticky, high-carbohydrate, nondomesticated plants resulted in extensive caries.
High caries prevalence in some Mesolithic foragers from Sicily and Portugal (Borgognini
Tarli et al., 1985, 1993; Frayer, 1988a; Lubell et al., 1994; Meiklejohn et al., 1988; but see
Cunha et al., 2007) contrasts sharply with low caries rates in other Mesolithic-period European
(e.g., Scandinavia: Alexandersen, 1988; Bennike & Alexandersen, 2007; Meiklejohn &
Zvelebil, 1991); Meiklejohn et al., 1988, 1997), South Asian (e.g., North India: Lukacs & Pal,
1993, 2003), and African (e.g., Nubia: Rose et al., 1993) populations. The high prevalence of
dental caries in Sicily and Portugal has been linked with the consumption of cariogenic
nonagricultural foods (e.g., honey) or sweet, sticky fruits (e.g., dates or figs) (Lubell et al.,
1994). Comparison of foragers (Mesolithic) and later farmers (Neolithic) from Portugal
revealed an increase in caries prevalence with the adoption of agriculture (Cunha et al., 2007).
Analyses of other hunter-gatherers point to the importance of additional factors that
explain unusually high caries prevalence. In South and Southeast Asia, quite high caries
prevalence in pre-Neolithic contexts points to dietary patterns that promoted caries in either
the absence of plant domestication, or at least only limited access to consumption of crops.
Lukacs (1990) documented relatively high caries prevalence (8.0%) in a South Asian Meso-
lithic population (Langhnaj site) in comparison with other contemporary series from the
region (most are approximately or much lower than 1%). Archaeological evidence indicates
close trade relationships between Langhnaj and agricultural groups, suggesting that cario-
genic foods may have been acquired through exchanges (Lukacs, 1990). In some living
foraging groups in central Africa, although called “hunter-gatherers,” significant consump-
tion of cariogenic plant foods acquired by trade from agricultural villagers results in appre-
ciable prevalences of caries (Walker & Hewlett, 1990; and see later). At Mac Bac, Northern
Vietnam, caries prevalence of 11% in adults is considerably higher than that of other
contemporary mid-Holocene populations in the region (Oxenham & Dommett, 2011). There
is no evidence for rice consumption, suggesting that other plant sources of carbohydrates

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.2 Dental caries 73

(e.g., taro) may have led to such a remarkably high prevalence of caries (compare with Tayles,
1999; Temple, 2007). Similarly, in other settings of late Pleistocene and early Holocene
populations where nondomesticated plant carbohydrates were consumed, dental caries preva-
lence is highly elevated, especially in comparison with most hunter-gatherers (Da-Gloria &
Larsen, 2014; Humphrey et al., 2014).
Changes in diet and subsistence technology had far-reaching implications for oral health in
some prehistoric foragers. Walker and Erlandson (1986) examined the link between dental
caries and dietary change on Santa Rosa Island, in the Santa Barbara Channel Island region of
southern California. Archaeological evidence indicates that exclusively foraging groups
occupied the island from c. 4000 to 400 yBP. For the first 1500 years of this timespan,
populations exploited predominantly terrestrial foods, primarily starchy roots and tubers.
For the remainder of the prehistoric period, diet became increasingly focused on marine
resources (see Chapter 2). A decrease in adult caries prevalence from 13.3% to 6.3% coincides
with this subsistence change, which appears to be linked to the reduction in use of plant
carbohydrates in later prehistory (Walker and Erlandson, 1986).
Other foraging groups show the opposite trend in caries prevalence. In Archaic-period
hunter-gatherers living in the Edwards Plateau of central Texas, there is an increase in
cariogenesis (Bement, 1994). This increase may reflect a reorientation of diet from a general-
ized foraging pattern involving a range of wild plants and animals to a diet focused more on
high-carbohydrate, cariogenic wild plants (Bement, 1994).

3.2.3 Sex differences in caries prevalence, dietary behavior, and


reproductive history
Comparisons of a wide range of archaeological populations from different times and settings
reveal a common pattern of greater caries prevalence in females than in males. In eastern
North America, regional studies point to this widespread difference, particularly in late
prehistoric and historic-era maize farmers or those with otherwise high carbohydrate, agri-
cultural diets (Behrend, 1978; Blakely, 1995; Bruwelheide et al., 2010; Hrdlička, 1916; Kestle,
1988; Lambert, 2002a; Larsen, 1983, 1998; Larsen et al., 1991; Mack et al., 2009; Milner,
1984; Newman & Snow, 1942; Okazaki et al., 2013; Patterson, 1984; Seidel, 1995). In the
Georgia Bight, adult females have significantly more carious teeth than adult males, both in
late prehistory (15.2% vs. 10.9%) and during the late mission period (41.9% vs. 36.3%) (Larsen
et al., 1991). Populations show similar differences from other regions of North America
(Burns, 1979; Danforth et al., 1997; Dickel et al., 1984; Hooton, 1930; Márquez Morfín
et al., 2002; Schmucker, 1985; Vega Lizama & Cucina, 2014; Whittington, 1989) and else-
where (Bennike, 1985; Da-Gloria & Larsen, 2014; Delgado-Darias et al., 2005; Domett, 2001;
Formicola, 1986 – 1987; Frayer, 1984, 1988a; Hemphill, 2008; Henneberg, 1996; Hillson,
1979; Lukacs, 1992, 1996, 2008; Lukacs et al., 1989; Meng et al., 2011; Morris, 1992;
Oxenham, 2006; Oxenham & Domett, 2011; Pietrusewsky & Ikehara-Quebral, 2006; Swärd-
stedt, 1966; Stewart, 1931; Temple, 2011; Temple & Larsen, 2007, 2013; Walker & Hewlett,
1990; Whittington, 1999; but see Barmes, 1962; Burns, 1979, 1982; Clarkson & Worthington,
1993; Klaus & Tam, 2010; Moore & Corbett, 1973; Pietrusewsky, 1988; Pietrusewsky &
Douglas, 2002a; Powell, 1988; Rowe, 1982; Sciulli & Oberly, 2002; Sutter, 1995; Watson
et al., 2010; Wells, 1980; White, 1994).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
74 Exposure to infectious pathogens

Differences in caries prevalence between males and females suggest that food consumption,
both in diet and behavior, may have been different between sexes, with males consuming
more animal sources of protein than do females and females consuming more plant carbohy-
drates than do males. This conclusion is consistent with differences documented in female and
male subsistence behavior in historic and recent agriculturalists and foragers. For example,
Southeastern North American Indian women were responsible for most plant gathering and
agricultural activities such as planting, harvesting, and food preparation. Men were respon-
sible for hunting as their primary subsistence task (Hudson, 1976; Swanton, 1942, 1946; Van
Doren, 1928). Among foragers living at Ngarulurutja, Australia, Hayden observed “that in
spite of rules about sharing, the persons who did the most hunting ate the most meat. It is
clear that the young men who actually caught the game consumed most of it” (1979:166) (see
other accounts of sex differences in diets in various traditional contexts by Hawkes et al.,
2001; Hayden, 1979; Hewlett et al., 1982; Holtzman, 2009; Lee, 1968; McArthur, 1960;
Meehan, 1977; O’Connell et al., 1992; Okazaki et al., 2013; Walker & Hewlett, 1990; Wood-
burn, 1968).
Sex differences in caries prevalence in archaeological samples are identified in populations
practicing agriculture to some degree. However, female and male foragers from the Santa
Barbara Channel Island region display differences in caries prevalence for much of prehistory
(Lambert & Walker, 1991; Walker & Erlandson, 1986). Ethnographic observations indicate a
distinctive sexual division of labor in which men hunted and fished and women collected
plants. Historic accounts indicate that men ate more of the game they hunted than did women.
Thus, it appears that greater consumption of plants by women – a factor related to their
subsistence responsibilities – is reflected in their greater caries prevalence. Similarly, con-
sumption of cariogenic plants (wild tubers and fruits) is hypothesized to explain the higher
prevalence of dental caries in females than in males in Paleoamericans from Lagoa Santa,
Brazil (Da-Gloria & Larsen, 2014). In this setting, while nearly all tooth types show greater
caries prevalence in females, the differences are especially pronounced in the second and third
molars.
Walker and Hewlett (1990) investigated dental caries in several groups of central African
pygmy foragers (Aka, Mbuti, Efe) and Bantu farmers in order to build support for a behavioral
interpretation of prevalence variation. Comparisons of subsistence patterns and food con-
sumption practices in these groups reveal clear intra- and inter-population differences in
caries prevalences that provide an important perspective on dietary behavior and cariogen-
esis. Aka and Mbuti foragers practice net-hunting, and the Efe foragers bow-hunting. The
sources of animal protein available to these groups vary, depending on the season (e.g., meat
from seasonally available animals), and is mostly obtained through hunting and collecting.
Manioc and other cultigens acquired through trade with Bantu agriculturalists form a signifi-
cant component of their diets. The relative consumption of meat to cultigens is highly
variable, ranging from a high proportion for the Mbuti (at least part of the year) to a much
lower proportion for the Efe. The Aka apparently consume similar amounts of meat as Mbuti
foragers as both are net-hunters. Based on comparisons of the amount of time spent hunting,
it appears that Mbuti and Aka foragers consume more meat than do Efe foragers. In addition
to meat and cultigens, honey is an important component of diet for part of the year. For Mbuti
foragers, this highly cariogenic food contributes nearly 80% of calories for a one-month
period (Ichikawa, 1981).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.2 Dental caries 75

Table 3.1 Sex differences in central African dental caries prevalences: Aka, Mbuti,
Efe, Bantu (Adapted from Walker & Hewlett, 1990: Table 2)
Teeth (n) % caries
Tribe and sex
Aka
Total (3099) 5.2
Males (1706) 4.2
Females (1393) 6.6
Mbuti
Total (1773) 6.0
Males (1048) 5.1
Females (753) 7.3
Efe
Total (277) 6.0
Males (277) 6.0
– –
Females
Bantu
Total (630) 8.1
Males (308) 9.1
Females (322) 7.1

Virtually all foods consumed by Bantu villagers are plant domesticates, including manioc,
maize, rice, peanuts, and plantains. The little amount of meat that is consumed –
approximately 2.5% of foods – is acquired from pygmy foragers. Ethnographic field observa-
tions indicate differences in food consumption between females and males in the foragers, but
not in the farmers. Diets of forager women contain more plants than those of forager men.
Aka men consume more meat than do women, which is acquired on the hunt prior to their
return to the home village; some of the choicest cuts of meat are shared among the men who
participated in the hunt.
Bantu farmers have the higher dental caries prevalence in comparison with foragers
(Table 3.1), which appears to reflect their greater consumption of plant carbohydrates.
Although foragers have relatively lower caries prevalence, the values, nevertheless, are
appreciable (5% – 6%: Table 3.1), which points to the significant component of plant carbo-
hydrates acquired from Bantu farmers via trading relationships. Caries prevalence is also
higher for pygmy women than men. These differences are related to food consumption
variation between male and female pygmy foragers. The frequency of between-meal eating
in pygmy women is an additional factor that provides partial explanation for sex differences.
In these foragers, men concentrate their eating into several large meals, and women snack
frequently during the day. Clinical evidence from Western populations indicates that snacking
between meals (especially carbohydrates) results in elevated caries rates (Burt et al., 1982,
1988; Gustafsson et al., 1954; Konig, 1970; Konig et al., 1969; Nizel, 1973; Rowe, 1982; Weiss
& Trihart, 1960; but see Rugg-Gunn et al., 1984).
Ethnographic documentation of dietary practices in native groups in South America
provides additional insight into differences between females and males in dental health

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
76 Exposure to infectious pathogens

(Walker et al., 1998). In the three groups studied – the Yanomami of Venezuela, the Yora of
southeastern Peru, and the Shiwiar (Achuar) of Ecuador – meat and fish provide a significant
part of the diet, and most carbohydrates are from plant crops, such as manioc and bananas.
All three groups have significant caries rates, in part due to the consumption of cultigens, but
also due to the reduction in their isolation and greater access to processed foods. This is an
unexpected finding because Yora and Shiwiar women spend many hours processing manioc
in their mouths for production of beer (chicha). For these two groups, women display a higher
frequency of carious teeth than men. Their relatively greater exposure to cariogenic manioc
explains these higher caries rates.
Given the finding of predominantly higher prevalence of dental caries in females than in
males in archaeological contexts discussed here and in living populations (Lukacs, 2008;
Lukacs & Thompson, 2008; Sauerwein, 1974), non-behavioral reasons relating to life history
factors may also contribute in important ways to this prevalence sex dichotomy. Permanent
teeth erupt slightly earlier in females than in males, exposing their teeth at an earlier age to
caries-promoting factors (Carlos & Gittlesohn, 1965; DePaola et al., 1982; Dunbar, 1969).
However, tooth eruption differences between males and females show either weak or no
correlation with dental caries prevalence (Moorrees, 1957; Toverud et al., 1952; Ziskin, 1926)
and the absolute temporal difference in tooth eruption is not significant (12 – 16 months).
Much more likely than tooth emergence timing as a cause for sex differences in caries
prevalence are factors relating to changes in the oral environment during pregnancy owing
especially to increased production of estrogens and progesterone (Laine, 2002). In this regard,
there is considerable evidence for increasing gingival inflammation in pregnant women
(Arafat, 1974; Laine, 2002; Loe, 1965; Loe & Silness, 1963; Maier & Orban, 1949; but see
Jonsson et al., 1988). At least one meta-analysis shows, however, that inflammation is
restricted to the gingiva, does not affect the tooth supporting tissue, especially the periodontal
ligament, and generally does not result in tooth exfoliation (Laine, 2002). In addition to these
changes, bacteria are altered in the oral environment, including an increased presence of
Streptococcus mutans and other flora associated with cariogenesis, and there is a lowering of
pH. However, most investigations of caries and pregnancy are largely cross-sectional or with
limited follow-up (Laine, 2002). Moreover, because dental caries is a process that unfolds over
several years or longer, it is difficult to determine a link between caries and pregnancy. The
record is also silent on the rate of caries during pregnancy versus before or after pregnancy.
Certainly, pregnancy influences the oral environment, which could result in increased cario-
genesis. The combined effects of more cariogenic bacteria, decreased pH, and the altered
buffer effect of saliva may contribute to increased dental caries in pregnant women. On the
other hand, the research tracking change in the oral environment shows that oral buffering
capacity fluctuates during pregnancy – decreasing, increasing, then decreasing – and increas-
ing dramatically after birth (Laine et al., 1988). Therefore, the idea that the oral environment is
simply more cariogenic in pregnancy does not take into account the full pattern of oral
buffering capacity during (or after) pregnancy.
The comprehensive investigation linking increased caries to life history in women by
Lukacs and coworkers (Lukacs, 2008; Lukacs and Thompson, 2008) and potential links with
increased fertility and birthrate in the Holocene, and the additive effects of a greater number
of pregnancies in agricultural settings presents a convincing scenario that may explain, in
part, the differences in dental caries prevalence of adult males and females globally.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.2 Dental caries 77

Bioarchaeologists need to include behavioral, biological, and archaeological contexts in


developing informed models for understanding this ubiquitous condition (Klaus & Tam,
2010; Lukacs, 2008, 2012; Temple, 2011). Many questions remain unresolved, however, but
can best be addressed by longitudinal studies (not possible with archaeological skeletal
samples). Such an investigation involving living populations will provide a better understand-
ing of timing and rate of dental caries before, during, and after pregnancy.

3.2.4 Status differences in dental caries


There is growing evidence to suggest that members of different social rankings consumed
dissimilar foods, leading to contrasting patterns of dental disease. High-status Edo-period
Japanese, Yin-Shang-period (Shang dynasty) Chinese, and Dynastic-era Egyptian adults have
a higher prevalence of caries (and other oral health indicators) than low-status adults (Leigh,
1934; Sakashita et al., 1997; Suzuki et al., 1967). A strong case is made in each of these
settings that the differences are related to dietary consistency: high-status adults in these
settings consume much softer, more refined foods than do low-status individuals. Softer food
consistency promotes the development of plaque and oral environments favorable to flora
that result in caries. The masticatory pattern presented in these settings also results in
different levels of masticatory loading and craniofacial robusticity (see Chapter 7).
Comparisons of dental caries prevalences of upper-class and lower-class adults from
Medieval Europe show contrasting patterns. At Westerhus, Sweden, upper-, middle-, and
lower-class individuals buried in and around a church show no differences in caries preva-
lence (Swärdstedt, 1966). The lack of class distinctions in dental health suggests that diets in
this area of Medieval Sweden were similar, regardless of social rank. In contrast, dental health
in a Medieval population from Zalavár, Hungary, clearly varied along social lines (Frayer,
1984). Individuals associated with the castle (upper class) have 6.4% carious teeth and
individuals from the chapel (lower class) have 12.1% carious teeth. Frayer (1984) argues that
these differences reflect greater consumption of animal protein in the upper class than in the
lower class. Similarly, individuals interred nearest the altar have lower caries prevalence than
those interred furthest from the altar at San Michele’s church in Trino Vercellese, Italy
(Girotti & Doro Garetto, 1999). Information on specifics of dietary variation in social classes
in Medieval Sweden, Hungary, and Italy are not available. Historical records for contemporary
populations from Britain indicate that upper-class landed gentry had greater access to animal
sources of protein, and their diet included a relatively smaller proportion of carbohydrates
than that consumed by lower-class peasantry (Wells, 1975). It seems highly likely that similar
variation in access to protein existed in these settings generally.
Some of the best evidence for the links between class structure and oral health in New
World settings is from the Classic period of the Maya (c. AD 250 – 900). In a number of centers,
elites show differences in caries prevalence in comparison with non-elite members of the
population. Elite from the Classic-period Maya centers at Copan, Honduras, and Lamanai,
Belize, exhibit lower caries prevalence than do non-elites (White, 1994; Whittington, 1999).
Similarly, in northern Petén, at Calakmul and nearby Dzibanché and Kohunlich, low-status
individuals from the combined sample express more carious lesions than do high-status
groups (Cucina & Tiesler, 2003). These findings indicate that high-status persons consumed
less maize than did low-status persons. The record of diet in this region shows greater access

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
78 Exposure to infectious pathogens

to animal sources of protein and greater diversity of diet in elite individuals (Cucina & Tiesler,
2003). The greater nutritional quality, including relatively lower carbohydrate (maize) con-
sumption in elites, explains their better oral health compared to non-elites.
On the north coast of Peru, there appears to be significant status-based variation in dental
caries in the Middle Sicán theocratic state (AD 900 – 1100) (Shimada et al., 2004; Klaus,
Wilbur et al., 2010). The Middle Sicán featured rigid social stratification with multiple
archaeological reflections of immense gulfs of power between the elite Sicán and the popula-
tion of Muchik commoners they ruled over. Dental caries prevalence is significantly lower
among the Middle Sicán elite, who probably consumed more protein resources and less
carbohydrates, as opposed to the Muchik commoners, whose higher caries prevalence indicate
consumption of relatively greater proportions of starchy cultigens in their diets. This pattern
of lower caries prevalence in elites is also strongly pronounced in complex societies in the
Atacama Oases of Chile. Like the Middle Sicán elite, adult elite from Solcor 3 and Coyo 3 sites,
especially males, express considerable differences (Hubbe et al., 2012). In this setting, the
better oral health in elites appears to be related to their having relatively greater access to
animal sources of protein.
Winkler and coworkers (2012) test the hypothesis that the relative position of skeletal
remains located closest to the ritual nucleus of the church (the altar) at the seventeenth-
century mission, Santa Catalina de Guale, Georgia, reflects higher social position in the
community, and hence, greater access to resources and better living circumstances. To test
this hypothesis, they undertook a location analysis of dental caries as an indicator of dietary
quality, nutritional inference, and generalized stress. Status rank was identified on the basis of
elaborateness of grave inclusions (e.g., presence of glass beads) and burial treatment (e.g.,
inclusion in a coffin). Teeth of individuals interred closest to the altar were significantly less
carious than those of individuals interred furthest from the altar. Their study suggests that
individuals buried closer to the altar had a higher-quality diet and greater access to resources
than those buried further from the altar. These results strongly suggest clear social distinctions
in health and well-being in this colonial setting of North America.
An argument for greater access to animal protein in explaining better dental health in high-
status members of human populations is supported by ethnographic documentation of status
differences in dental health in African foragers. Walker and Hewlett (1990) found that high-
status pygmy leaders have fewer carious lesions than low-status non-leaders. They suggest
that greater access to meat by leaders (from gifts and tribute), combined with reduced
consumption of carbohyrate-rich plants, best explains the discrepancy between social ranks.

3.3 Periodontal disease (periodontitis) and tooth loss


3.3.1 Description and etiology
The accumulations of bacteria (plaque) on teeth are closely linked to an inflammation of the
tissues surrounding teeth, a condition known as periodontal disease (Loesche & Grossman,
2001). Periodontal disease is readily visible in soft tissue – the gums appear swollen and red.
The inflammation results in loss of the collagen attachment of the tooth to the supporting
bone. Eventually, the supporting alveolar bone is so depleted that it results in exfoliation of
the teeth. Another potential irritant to the soft tissues surrounding alveolar bone is excessive

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.3 Periodontal disease (periodontitis) and tooth loss 79

(a)

Figure 3.3a Antemortem tooth loss of mandibular dentition; Ochsenfurt, Germany.


(Photograph and copyright by Leslie L. Williams; reproduced with permission.) (A black and
white version of this figure will appear in some formats. For the color version, please refer
to the plate section.)

masticatory loading of the jaws and teeth. These mechanical demands are due to either
consumption of hard-textured foods or excessive extramasticatory practices, such as process-
ing of animal hides (Bement, 1994; Clarke & Hirsch, 1991; Marks et al., 1988; Molnar, 1972;
Pedersen & Jakobsen, 1989). Extreme mechanical demands on anterior teeth in northern
latitude populations result in severe tooth wear, pulp exposure, and resorption and shortening
of tooth roots. These factors contribute to tooth loss (Pedersen & Jakobsen, 1989).
Periodontal disease or periodontitis is a very common disease in industrialized countries
today (about a third of the adult population in the United States: Buhlin et al., 2003). It is
generally characterized by a loss of alveolar bone, represented either by a horizontal lowering
of the alveolar crest relative to the neck of the tooth (cementoenamel junction) or as pockets
of bone rarefaction (Hildebolt & Molnar, 1991) (Figure 3.3a). In the clinical setting, horizontal
bone loss is determined by the distance between the alveolar crest (interproximal bone) and
the cementoenamel junction. This dimension may not always represent an appropriate
indicator of periodontal disease, especially in anthropological populations where continuous
eruption during adulthood occurs in response to occlusal wear (Clarke, 1993; Clarke & Hirsch,
1991; Whittaker et al., 1985). Clarke and Hirsch suggest that periodontal disease in skeletal
remains is identified when “the crestal margin of bone undergoes loss of the surface cortical
bone, exposing the porous cancellous structure of the supporting bone, usually with an
accompanying change of the contour of the crest” (1991:241). Thus, exposure of tooth roots
as representing periodontal disease or simply compensatory continuous eruption owing to

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
80 Exposure to infectious pathogens

(b)

Figure 3.3b Edentulous individual; Santa Catalina de Guale Santa Maria, Amelia Island,
Florida. (From Larsen, 1994; photograph by Mark C. Griffin.) (A black and white version of
this figure will appear in some formats. For the color version, please refer to the plate
section.)

wear and mechanical loading must be based on the quality of the alveolus – rough, irregular,
and porotic in periodontal disease, but thin, knife-edged, and dense in continuous eruption
(Ogden, 2008).
Once teeth are gone (Figure 3.3b), the soft tissue heals, and the alveolus remodels com-
pletely. Although the progression of periodontal disease is well documented in human
populations, ancient and modern, it remains an etiological conundrum. There is general
agreement that bacteria – perhaps as many as 40 different taxa – may be involved in the onset
and progression of the disease (Drake et al., 1993; Enwonwu, 1995; Li et al., 2000; Slots, 2004;
van Winkelhoff and Slots, 1999). Other important influencing factors include poor oral
hygiene, cariogenesis, malocclusion, nutritional status, and to a lesser extent, pregnancy,
puberty, and psychological stress (Hildebolt & Molnar, 1991). Some individuals appear to be
more susceptible, due in part to the presence of inflammatory gene polymorphisms (cytokines)
that determine greater presence of inflammatory serums (e.g., C-reactive protein: D’Aiuto

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.3 Periodontal disease (periodontitis) and tooth loss 81

et al., 2004). Those individuals with genotypes specific for these cytokines appear to have
elevated levels of systemic inflammation with periodontal infection exacerbating the already
overloaded inflammatory burden. However, these genotypes are rare, and the general levels of
elevated periodontitis in human populations are largely due to environmental factors, espe-
cially diet and oral hygiene.

3.3.2 Temporal trends: foragers, farmers, and industrialized populations


Many of the foregoing populations with elevated prevalence of dental caries also exhibit high
frequencies of tooth loss. The high prevalence of dental caries in populations with tooth loss
indicates an association between the two conditions. It is difficult to identify cause of
antemortem tooth loss in most instances. Given the similarities between clinical populations
and archaeological remains in location and pattern of tooth loss – usually commencing with
the posterior mandibular dentition – tooth loss in archaeological contexts is linked closely to
periodontal disease.
Few workers have reported systematically on tooth loss prevalence in past human popula-
tions. The paucity of data reflects a lack of consensus on etiology (Clarke & Hirsch, 1991;
Hildebolt & Molnar, 1991), as well as the poor representation of intact alveolar bone in many
archaeological remains. The available record indicates that tooth loss has an ancient history.
Various early hominin remains show evidence of alveolar bone resorption, including Homo
erectus (e.g., Mauer) and late archaic Homo sapiens (e.g., Krapina, La Chapelle-aux-Saints)
(Hildebolt & Molnar, 1991; Hillson, 2005; Trinkaus, 1985; Wells, 1975). Significant tooth loss,
however, is largely limited to Holocene populations.
As with dental caries, recent populations undergoing shifts from traditional to Western
processed diets experience a marked increase in periodontal disease and tooth loss (Barmes,
1977; Basu & Dutta, 1963; Clarke et al., 1986; Cutress et al., 1982; Heithersay, 1959; Homan,
1977; Mayhall, 1977; Moorrees, 1957; Tal, 1985; Walker et al., 1998; and many others).
Similarly, archaeological populations showing evidence of elevated or increasing levels of
consumption of plant carbohydrates or processed foods have high or accelerated rates of
periodontal disease and tooth loss (Alfonso et al., 2007; Bennike & Alexandersen, 2007; Blau,
2007; Delgado-Darias et al., 2006; Douglas & Pietrusewsky, 2007; Frayer, 1984; Kerr, 1991,
1998; Klaus & Tam, 2010; Larsen, Craig et al., 1995; Nelson et al., 1999; Owsley et al., 1987;
Patterson, 1984; Sakashita et al., 1997; Sledzik & Moore-Jansen, 1991; Smith & Horwitz,
2007). In contrast, many foragers with diets dominated by animal protein have low preva-
lences (Costa, 1980; Scott et al., 1991).
The shift from foraging to farming was accompanied by an increase in periodontal disease
and tooth loss. In the Eastern Woodlands of North America, tooth loss increases dramatically
with the shift to farming (Cassidy, 1984; Cook, 1984; Patterson, 1984; Smith, 1982). In
prehistoric Ontario, for example, molar loss increases from a low of 12.0% (LeVesconte
Mound) to 37.1% (Glen Williams site) in prehistoric foragers and farmers, respectively
(Patterson, 1984). This and other populations show a higher prevalence of tooth loss in
posterior teeth (especially molars) and in older adults. The cause for tooth loss in Ontario
appears to be different in comparing foragers and farmers. In foragers, tooth loss is due to
pulp exposure from severe occlusal wear, and in farmers, tooth loss is due to periodontal
disease and dental caries.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
82 Exposure to infectious pathogens

45

40 Late pre-Hispanic period


Postcontact period
35

30
Prevalence AMTL (%)

25

20

15

10

0
1 2 3 4 5 6
Age Class
Figure 3.4 Comparison among age classes of posterior antemortem tooth loss (AMTL) in
late pre-Hispanic and postcontact periods, Lambayeque Valley Complex, Peru. Age classes
are defined as 0–4.9 yr (1), 5.0–14.9 yr (2), 15.0–24.9 yr (3), 25.0–34.9 yr (4), 35.0–44.9 yr
(5), and 45þ yr (6). (Adapted from Klaus & Tam, 2010; reproduced with permission of
authors and John Wiley & Sons, Inc.)

Other settings in Europe, Asia, and Africa mirror the trends documented in North America.
Seventeenth-century British mandibles exhibit lower crestal bone than Anglo-Saxon (sixth
century AD) mandibles (Lavelle and Moore, 1969). Increasing prevalence of periodontal
disease in this setting is due to the greater consumption of softer foods in the later period,
mostly arising from improvements in milling flour for bread and increased consumption of
sugar and refined carbohydrates generally. This trend closely parallels rapid increases in
dental caries in Britain for the same period (Hardwick, 1960; Moore & Corbett, 1975).
On the Indian subcontinent, Lukacs (1992) documented tooth loss in an extensive series of
human remains from various sites dating from the Mesolithic through the Iron Age, a
temporal framework representing the transition from foraging to intensive agriculture. The
later agricultural populations consumed a variety of domesticated plants (wheat, barley, peas,
sessamum). Comparisons of the earlier and later populations reveal an increase in tooth loss
resulting from an increase in consumption of plant carbohydrates (Lukacs, 1992). Similar
patterns of increased loss are found with the adoption and later intensification of agriculture
in Southeast Asia (Douglas, 2006) and South America (Alfonso et al., 2007; Klaus & Tam,
2010) (Figure 3.4).
Declining oral health has been identified in the Nile Valley in comparing dentitions from
four successive periods: Mesolithic (c. 15 000 yBP), Meroitic (350 BC – AD 350), X-group

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.3 Periodontal disease (periodontitis) and tooth loss 83

(AD 350 – 550), and Christian (AD 550 – 1300) (Rose et al., 1993). The Mesolithic group is
characterized by a foraging economy based largely on seed gathering and hunting. Popula-
tions in the three later periods were irrigation farmers who grew cereal grains. All samples
show evidence of tooth loss. However, the Mesolithic group possessed only two teeth lost prior
to death (from 400 sockets representing 39 individuals). Later, the prevalence steadily rises
from 9.9% (Meroitic) to 34.6% (Christian). Rose and coworkers (1993) suggest that tooth loss,
although multifactorial, was due to extensive caries and excessive occlusal wear, which
predisposed teeth to decay and eventual loss. Some of the tooth loss may be due to the use
of stone grinding implements and incorporation of grit into food. The resulting wear and pulp
exposure contributes to periodontal disease and tooth loss in other areas of the Nile Valley
(Marion, 1996).
Prevalence of periodontal disease may be overestimated in archaeological samples. Hyper-
eruption of teeth (and eventual loss) may represent a normal aging process bearing little
relation to pathological processes (Clarke & Hirsch, 1991; Whittaker, 1993; Whittaker et al.,
1985). However, the important role of the dentition in food consumption argues that tooth
loss is an indicator of compromised health.

3.3.3 Sex and status differences


Sex differences in the prevalence of tooth loss do not present a consistent pattern in
archaeological settings. However, there is some evidence to suggest that males have a greater
risk of developing periodontal disease than females. That is, the record for both past and living
societies reveals a tendency, albeit inconsistent, for higher prevalence and severity in males
(Costa, 1982; DeWitte, 2012; Jordan et al., 2011; Lukacs, 2007b; Pihlstrom, 2001; Šlaus, 2000;
but see Al-Shammery et al., 1998). Cultural behavior is an important influence in understand-
ing this variation, however. In South American native populations recently coming into
contact with Western society, females have a higher prevalence of antemortem tooth loss
than males (Klaus & Tam, 2010; Walker et al., 1998). Walker and coworkers found that women
of child-bearing age use sap to extract diseased (carious) teeth, while on the north coast of
Peru, female antemortem tooth loss in Colonial Mórrope is associated with the installation of
a Western-style food-producing economy and changing manner of oral hygiene.
Very few studies of archaeological dentitions report female and male prevalence data, thus
preventing observations of sexual dimorphism in tooth loss. Populations from Bronze Age
Harappa, India (Lukacs, 1992) and Mesolithic Portugal (Frayer, 1988a) – both settings
involved consumption of plant carbohydrates and other cariogenic foods – as with dental
caries, have a higher prevalence of tooth loss in females than in males. Similarly, females in
some settings in southeast Asia and western Europe have higher prevalence of antemortem loss
than males (Domett, 2001; Oxenham, 2006; Prowse, 2011). While there is a tendency for these
populations to express co-morbidities of high caries rates and antemortem tooth loss, the
association is weak in a number of settings, indicating that antemortem loss is not necessarily
caused by caries. Nonetheless, although not systematically analyzed in bioarchaeological
research, the common presence of both conditions in individuals and populations indicates
the common outcome of poor diets involving consumption of soft-textured carbohydrates.
With respect to the anterior dentition, differences are quite pronounced in the comparison of
female and male anterior tooth loss: females have much greater anterior tooth loss than males.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
84 Exposure to infectious pathogens

25
Male

Female
% Antemortem Tooth Loss 20

15

10

0
I1 I2 C P3 P4 M1 M2 M3 I1 I2 C P3 P4 M1 M2 M3
Maxilla Mandible

Figure 3.5 Percentage of antemortem tooth loss in males and females; Canary Island
archipelago. (Lukacs, 2007b; reproduced with permission of author and John Wiley & Sons,
Ltd.)

In the Ipiutak series, males have a very low loss of incisors (5.3%); females display a high loss
of incisors (19.4%). Hrdlička (1940a) argued that tooth loss in this region was related to the
practice of tooth ablation whereby the anterior teeth were intentionally removed (Cook,
1981b; and compare with Temple et al., 2011). Re-examination of cultural practices in these
groups suggests that greater anterior tooth loss in adult females is related to the excessive use
of front teeth in extramasticatory activities (Costa, 1980). Women in this setting engage in
behaviors that place excessive mechanical demands on the anterior teeth, such as hide
chewing. The cumulative trauma and excessive dental wear result in early loss of incisors
and canines. Thus, tooth loss differences between men and women can clearly be related to
use of the dentition in extramasticatory behaviors. The links with trauma are also well
illustrated in sex difference in antemortem tooth loss in the Canary Islands (Lukacs, 2007b),
where males generally have greater antemortem loss than females (Figure 3.5). In this setting,
antemortem losses are due to multiple factors, including accidents from walking in difficult
terrain and ritual and non-ritual forms of interpersonal violence involving mostly males. This
body of work demonstrates the complex nature of antemortem tooth loss relating to a range
of factors. The record globally, past and present, is largely explained by oral health patterns
influenced by diet, but behaviors not involving food and its mastication are important.
The influence of status and social rank on tooth loss in human populations has rarely been
systematically assessed in archaeological remains. In the Medieval series from Zalavár,
Hungary, tooth loss is considerably higher in low-status individuals buried in the chapel
(39.4%) than in high-status individuals buried in the castle (9.1%) (Frayer, 1984). The
difference between status groups is especially striking in adult males: 48.4% of teeth were

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.3 Periodontal disease (periodontitis) and tooth loss 85

missing in the low-status group and only 5.2% were missing in the high-status group.
Females show prevalences of 32.1% and 6.7% for low- and high-status groups, respectively.
As with the dental caries prevalences from this series, sex differences in tooth loss indicate
dental health variation between men and women that is strongly influenced by sexual
dimorphism in diet and food consumption practices. In contrast, high-status persons from
the aforementioned Calakmul, Dzibanché, and Kohunlich in the northern Petén have higher
antemortem tooth loss than low-status persons (Cucina & Tiesler, 2003). In this setting, high-
status individuals express greater presence of calculus, which contributes to irritation to and
inflammation of gingival and periodontal tissues and increased risk of tooth loss. Overall, a
more refined, softer diet contributed to worse oral health in elite persons. A similar pattern is
present in Shang dynasty China where reduced tooth wear suggests a softer, more processed
diet and an oral environment more conducive to poor oral health in the comparison of
“citizens” and “slaves” (Sakashita et al., 1997).

3.3.4 Implications of poor oral health for general health: the mouth as a window
to systemic health
A wide range of clinical and epidemiological investigations have examined the relationship
between poor oral environment and general health, especially with regard to adult mortality
and chronic, systemic health conditions such as cardiovascular disease, respiratory infections,
and other diseases (Ajwani et al., 2003; Buhlin et al., 2003; DeStefano et al., 1993; Hollister
and Weintraub, 1993; Irwin et al., 2008; Johnson et al., 2006; Joshipura et al., 1996; Koren
et al., 2011; Li et al., 2000; Pihlstrom et al., 2005; Slots, 2004; Williams et al., 2008).
Periodontal disease and cardiovascular disease show the clearest association, with perio-
dontitis associated with increased risk of coronary heart disease. The mechanism is unclear
and confounding factors such as age, diet, sex, and smoking history are problematic in the
design of research and specific links. It may be the case that periodontal disease is the root
cause of systemic infection and the attendant inflammatory response, which may in turn
promote an inflammatory response in cardiovascular tissues. Nonetheless, the oral environ-
ment is a clear target for developing an understanding of adult systemic conditions and it is a
risk factor for increased mortality linked to cardiovascular conditions, pulmonary disease,
other evidence of morbidity, and increased mortality. In the last couple of decades, about 90%
of adults globally express periodontal disease. It is clear, therefore, that reconstructions of oral
health are crucial for understanding general health (DeWitte & Bekvalac, 2011; Pihlstrom
et al., 2005; Slots, 2004).
The links between inflammation associated with oral health and elevated mortality from
cardiovascular disease in living populations suggest the possibility of links between oral
health status and mortality in past populations. Indeed, the association between poor oral
health and increased risk of death (based on mortality analysis) is clearly illustrated in a large
series of remains from the medieval St. Mary Graces cemetery in London (DeWitte & Bekvalac,
2010). In this setting, there is a statistically significant relationship between oral health and
mortality: individuals with dental caries and periodontal disease are at greater risk of dying
than those without the conditions. Similarly, individuals having periodontal disease were far
more likely to have periosteal reactions than individuals not having periodontal disease
(DeWitte & Bekvalac, 2011). Historical sources show that regardless of status, individuals

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
86 Exposure to infectious pathogens

living in London at the time were consuming, as the largest part of their diet, wheat, barley,
and rye in various forms (e.g., bread, ale, cereal mixtures). This, combined with poor oral
hygiene, explains the elevated levels of oral disease in this setting. Those who enjoyed better
diets (especially including animal sources of protein) almost certainly had better oral and
systemic health, and hence, lower risk of death (DeWitte & Bekvalac, 2010, 2011).

3.4 Nonspecific infection and disruption


3.4.1 Periostitis and osteomyelitis
Skeletal lesions of infectious origin represent a continuum, initially involving the periosteum,
followed in severity by involvement of cortical bone generally, and at the extreme end,
extension of infection into the medullary cavity. Periostitis (or periosteal reaction) is the least
severe, and osteitis and osteomyelitis are the most severe lesions. Periostitis represents a basic
inflammatory response resulting in new bone formation. It may result from bacterial infec-
tion, but also from traumatic injury or virtually anything that disrupts the periosteum and
stimulates new bone formation (Weston, 2012). Infection or injury initiates an inflammatory
response that elicits bone production by stimulating the osteoblasts lining the subperiosteal
membrane (Eyre-Brook, 1984; Simpson, 1985). Specifically, lesions form following compres-
sion and stretching of blood vessels by pus, neoplasm, trauma, blood, granulation tissues, and
the follow-up response of the bone as a healing mechanism (see Weston, 2008, for discussion).
The resulting lesions are characterized as osseous plaques with demarcated margins or

Figure 3.6 Periosteal reaction on mid-diaphysis of adult tibia; Ochsenfurt, Germany.


(Photograph by Leslie L. Williams.) (A black and white version of this figure will appear in
some formats. For the color version, please refer to the plate section.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.4 Nonspecific infection and disruption 87

Figure 3.7 Osteomyelitis involving diaphysis of adult tibia and fibulae; King site, Georgia.
(Photograph by Clark Spencer Larsen.)

irregular elevations of bone surfaces (Figure 3.6). The skeletal tissue in the unhealed form is
loosely organized new bone. In the healed form, the skeletal tissue is incorporated into the
normal cortical bone and the surface is often smooth, undulating, and somewhat inflated. The
lesions can be highly localized, being limited to single skeletal elements, but they may also
involve multiple elements if the infection is widespread or systemic. Osteitis is usually not
identifiable without radiological observation. As most paleopathological studies do not
involve radiological analysis, osteitis is not discussed here.
Osteomyelitis involves exuberant proliferation of both endosteal and periosteal bone
surfaces (Figure 3.7), the former of which results in the restriction in diameter of the medullary
cavity. Most pyogenic (pus-producing) osteomyelitis is caused by the microorganism
Staphylococcus aureus (Aufderheide & Rodríguez-Martín, 1998; Rosenberg, 1994). The arter-
ial system of the bone is the typical route of its transport. Direct infection from a bone fracture
and break in the skin is also linked with osteomyelitis. The infection site on the bone is
sometimes associated with sinuses or holes (cloacae) for exudate or pus drainage. Chronic
osteomyelitis can occur over a period of many years due to the presence of localized infection
foci that occasionally reappear, sometimes in response to systemic or localized stress.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
88 Exposure to infectious pathogens

Infection resulting in periostitis is almost never fatal, as it is usually restricted to a localized


area of a single bone. Infection leading to osteomyelitis can result in death if the infection
spreads via the circulatory system to vital organs. As with the periosteal reaction, the bone
tissue associated with osteomyelitis has a woven, porous appearance while infection is active.
If healed, the bone is dense and becomes integrated with the normal underlying cortical tissue
(Ortner, 2003). Osteomyelitis is far less prevalent in archaeological series than periostitis
(Larsen, 1982; Powell, 1988).
Although periosteal reactions and osteomyelitis are generally nonspecific to a variety of
pathogens, their documentation has proven highly useful for assessing levels and patterns of
community health (various in Cohen & Armelagos, 1984; Cohen & Crane-Kramer, 2007;
Steckel and Rose, 2002). It is important to emphasize that periosteal reactions provide a rather
incomplete and nondiagnostic picture of a population’s disease experience (Cook, 2007),
largely because they can be caused by multiple factors, related to either infection or trauma.
I suspect that the association documented between elevated prevalence of periosteal reactions
seen in many archaeological contexts involving sedentism and population concentration and/
or poor living circumstances indicates an infectious origin (and see following discussion).
Moreover, the clear link with other inflammatory processes in the body suggests an infectious
component to periosteal lesions (DeWitte & Bekvalac, 2011). Thus, while periosteal reactions
should not be equated with infection (and see Weston, 2012), I regard the patterns docu-
mented in archaeological contexts as largely representing the association between living
circumstances, health, and bony responses involving infectious disease. In my view, it is an
important stress indicator having a nonspecific derivation. It may be the case that the trend
for increased periostitis was caused by a rise in injuries, especially involving the lower leg.
However, as discussed in the following section, the increase in conditions and circumstances
that promote infection, especially declining mobility, increased sedentism, and population
aggregation, suggest that the elevated prevalence of periostitis in later prehistory is more
likely related to pathological conditions and not traumatic injury.

3.4.2 Temporal trends: foragers and farmers


Periosteal reactions have been documented in many archaeological skeletal samples, but only
recently have biological anthropologists systematically collected data on prevalence in order
to make inter-population comparisons. These recent comparisons demonstrate important
prevalence trends in specific settings, especially in populations making the transition from
foraging to farming or in the comparison of foragers, farmers, and other lifeways.
Generally, populations undergoing adaptive shifts from foraging to part-time or intensive
farming show an increase in the prevalence of periostitis. Bioarchaeological analyses of
periosteal reactions reveal that the tibia is the most commonly affected bone (Cunha et al.,
2007; Gold, 2004; Hodges, 1989; Hutchinson & Larsen, 1995; Lambert, 1993; Larsen & Harn,
1994; Martin et al., 1991; Milner, 1991; Perzigian et al., 1984; Powell, 1988; Suzuki, 1991;
Ubelaker & Jones, 2003; Webb, 1995; Weston, 2012; and others). It is unclear why the rates
for the tibia are so much higher than those for other skeletal elements. Yet periosteal reactions
associated with syphilis show a penchant for the anterior tibia, the cranial vault, and the
superior aspect of the clavicular diaphysis, perhaps because these elements are not surrounded
by large amounts of soft tissue, and therefore, have slightly cooler temperatures, rendering

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.4 Nonspecific infection and disruption 89

them more susceptible to infection (Ortner, 2003). Additionally, the anterior and lateral
aspects of the tibial diaphysis have the largest, and perhaps the most vascular and physiolo-
gically inactive surfaces in the skeleton, which may also lead to bacterial colonization and
infection (Martin et al., 1991). Circulatory flow is generally slower in the lower legs because of
gravity, enhancing the potential for bacterial colonization (Kumar et al., 2010). The anatom-
ical location – the lower end of the legs – also exposes the anterior tibia to trauma against
which little protection is offered by soft tissue. Subcutaneous and subperiosteal bruises from
trauma promote periosteal reactions. In addition, these injuries also promote bacterial prolif-
eration through release of blood and intracellular fluids from ruptured cells and vessels,
resulting in a bony reaction.
In the Eastern Woodlands of North America, several studies show important links between
subsistence, settlement pattern, and community health. In the Tennessee River valley, non-
specific periosteal lesions increase in farmers relative to earlier foragers (Danforth et al., 2007).
Similarly, in the Dickson Mounds site in the Illinois River valley, Lallo and coworkers
(Goodman, Lallo et al., 1984; Lallo et al., 1978; Lallo & Rose, 1979) compared infection
prevalence in two populations, including an earlier Woodland (called Mississippian Accultur-
ated Late Woodland) group (AD 1050 – 1200) and a later Mississippian (Middle Mississippian)
(AD 1200 – 1300) group. The later period saw an increase in consumption of maize and
reduction in animal sources of protein. There appears to have been a consolidation of
residential units into larger population aggregates at the same time. Overall, settlement
involved a marked increase in population density, decrease in mobility, and conditions that
generally provide the context for increased infection.
There is a dramatic increase in frequency and severity of infectious lesions (periostitis and
osteomyelitis) that coincides with these changes in population settlement at Dickson Mounds,
Illinois. The prevalence doubled in the later period, from 30.8% to 67.4% of individuals
affected. For the tibia, the prevalence increased from 26% to 84%, affecting all age groups
and both sexes. Comparisons of severity (from slight to severe involvement of the periosteum)
for the tibia showed that most of the infection was slight in the Woodland group whereas most
was moderate to severe in the Mississippian group.
This increase in prevalence and severity of infection could be explained in a number of
ways. The introduction of new chronic infectious disease(s) may have occurred in later
prehistory. Lallo and coworkers argue that the increase in infectious lesions in the thirteenth
century AD is possibly tied to several interlinking factors. They highlight the role of decline in
nutritional quality and sedentism in later prehistory, which likely placed populations at
increased risk of infection by decreasing the resistance to disease and by providing lower
quality of living circumstances.
Archaeological evidence suggests that there was an increase in trade networks and long-
distance social contact in the later prehistoric period. These contacts may have provided a
means for introducing new pathogens or disease vectors, or both, thus increasing the
prevalence of infectious disease. The effects of population size increase and sedentism are
well understood in infectious disease ecology and epidemiology. By increasing the size and
density of settlements, the host and pathogen are placed side-by-side in a long-term relation-
ship that may form the basis of chronic infection. The number of potential hosts is increased,
thus providing a permanent reservoir for certain infectious agents. The closer contact in a
more densely occupied settlement, coupled with the ill effects of poor sanitation resulting

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
90 Exposure to infectious pathogens

from permanent occupancy of a setting, results in faster and more proficient disease trans-
mission (Armelagos, 1990; Armelagos & Dewey, 1970; Lambert, 1993; Lallo et al., 1978;
Larsen, 1982). Thus, the overall picture is one of declining community health as population
concentration increases. Moreover, the negative impact of infection on the population is
indicated by depressed survivorship for those individuals with skeletal lesions. For example,
Dickson adults (>20 years of age) with severe infections had an average age-at-death of 29.3
years, which is well below the mean age-at-death for the adult population overall (33.5 years)
(Lallo et al., 1978). This finding is consistent with the notion that individuals with elevated
stress earlier in life are predisposed to earlier death in adulthood (see Chapter 2).
Other settings from the Eastern Woodlands show results that are generally consistent with
changes observed in the Dickson Mounds populations. Comparisons of prehistoric foragers
(pre-AD 1150), prehistoric maize farmers (AD 1150 – 1550), early mission (AD 1607 – 1680),
and late mission-period intensive maize agriculturalists (AD 1686 – 1702) in the Georgia Bight
show clear temporal trends in prevalence in relation to dietary and lifeway changes (Larsen
and Harn, 1994; Larsen, Crosby et al., 2002; Larsen, unpublished; and see earlier for a more
detailed description of the samples).
Unlike the Dickson Mounds study, these analyses focus on comparisons of populations
consuming exclusively wild plants and animals with populations utilizing maize to varying
degrees. For most of the time prior to the arrival of Europeans in the sixteenth century, native
groups were nonsedentary foragers who obtained most foods from a combination of hunting,
gathering, and fishing. Archaeological and isotopic documentation of subsistence economy
indicates that marine resources from estuarine and marine contexts provided most of the
protein in native diet (Hutchinson et al., 1998; Larsen, Schoeninger et al., 1992; Reitz, 1988,
1990). During the twelfth century AD, maize rapidly took on an increasingly important
dietary role. With the arrival of Europeans and the establishment of Catholic missions, there
was a subsistence reorientation whereby maize dominated diet. Coupled with these dietary
changes were alterations in population size, density, and sedentism. Prehistoric foragers in the
region appear to have been sparsely settled and highly mobile. Prehistoric farmers, however,
were living in larger, more densely occupied villages, probably for longer periods of time
(Larsen, 1982). In the mission period, populations were coerced into living in and around
mission settlements. Although population size reduced dramatically during the contact
period, the settlements were permanent and villages were crowded.
Prevalence comparisons for periosteal reactions in adult tibiae indicate that there is an
increase in frequency prior to contact from 9.5% to 19.8%. In the early mission population,
the prevalence declined slightly to 15.4%, but then greatly increased to 59.3% in the late
mission period. These general increases affected both adult males and females. The moderate
increase in the precontact populations is well under that observed at Dickson Mounds. In all
likelihood, the relatively lower prevalence in the Georgia Bight reflects considerably smaller
population size and less of a commitment to maize agriculture in the later prehistoric period in
comparison with interior Mississippian populations (e.g., Dickson Mounds). The marked
increase in the late mission period is probably tied to the relocation and increased concen-
tration of native populations around mission centers and introduction of new diseases,
possibly including venereal syphilis. The change in settlement provided conditions conducive
to the maintenance and spread of chronic infectious diseases and other factors that lead to an
increase in bone lesions (Larsen, Crosby et al., 2002; Larsen & Harn, 1994; Larsen et al., 2007).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.4 Nonspecific infection and disruption 91

The effects of increased infection rates would likely have been exacerbated by the increase in
dietary emphasis on nutritionally poor foods, especially maize.
There is also a synergy between infection and malnutrition (Keusch & Farthing, 1986;
Scrimshaw et al., 1968) – malnourished individuals are less resistant to infectious pathogens
and are rendered more susceptible to infectious disease; conversely, infection worsens nutri-
tional status. In understanding the increase in infection in these archaeological (and other)
settings, the synergy with nutrition is critical. The consequences of infection and nutrition are
worse than either factor acting alone. Individuals experiencing infection exhibit higher basal
metabolic rates, which are accompanied by fever and the body’s increased demand for protein
and other nutrients necessary for the production of antibodies that fight the infection.
Therefore, in the setting of reduced nutritional quality, first in the late prehistoric and then
in the mission context, the ability to mitigate infection would likely have been hampered by a
reduction in dietary quality. Thus, infection increased in the mission period, in large part due
to the compromised health linked to poorer nutrition and population crowding.
These studies provide strong support for the epidemiological model that an increase in
population size and density often contributes to decline in community health, at least as it is
measured by prevalence of skeletal lesions. This general pattern of increase in infections with
the transition from foraging to farming or agricultural intensification is also documented in a
variety of other areas of the Eastern Woodlands undergoing the shift from foraging to farming
(Cassidy, 1984; Cook, 1984; Danforth et al., 2007; Garner, 1991; Hoyme & Bass, 1962;
Katzenberg, 1992; Pfeiffer & Fairgrieve, 1994; Rose et al., 1984, Rose, Anton et al., 1991) or
in single-component late prehistoric settings in the Eastern Woodlands (Boyd, 1986; Eisen-
berg, 1986, 1991a, 1991b; Gold, 2004; Magennis, 1986; Milner, 1982, 1991, 1992; Milner &
Smith, 1990; Powell, 1986, 1988, 1989). Increase in infection prevalence is also well docu-
mented in the American Southwest (Martin et al., 1991; Stodder, 1994; Stodder & Martin,
1992; Stodder et al., 2002), Mesoamerica (Hodges, 1989; Márquez Morfín & Storey, 2007;
Márquez Morfín et al., 2002; Norr, 1984; Saul, 1972; Storey, 1992a), South America (Klaus &
Tam, 2009; Pechenkina, Vrandenburg et al., 2007; Ubelaker, 1984, 1994), and in a few
regions of the Old World (e.g., Japan: Suzuki, 1991; Europe: Cunha et al., 2007; Grauer,
1991; Roberts & Cox, 2003, 2007).
This is not to say that increased population aggregation in farming societies invariably led
to the same levels of bone infection prevalence globally or even within broad regions. For
example, evidence of generalized infection and other health indicators are low in a number of
settings in Southeast Asia (Clark et al., 2014; Domett & Tayles, 2007; Douglas & Pietrusewsky,
2007). Community health in late prehistoric populations in eastern North America was highly
variable (Milner, 1991). Some late prehistoric agriculturalists had very high prevalences of
bone infections (Eisenberg, 1986; Lallo et al., 1978; Milner & Smith, 1990), while other groups
had somewhat lower prevalences (Larsen, 1984; Larsen, Crosby et al., 2002; Larsen & Harn,
1994; Larsen et al., 2007; Milner, 1991). Some of this variability is undoubtedly due to inter-
observer differences in recording methods. The variable pattern of infection prevalence also
points to a high degree of diversity between human groups occupying very different land-
scapes and physiographic zones, ranging from highly fertile river bottoms (Lallo et al., 1978)
to marginal uplands (Eisenberg, 1986) or coastal regions (Larsen, 1982). Detailed analysis of
population trends indicates that population histories fluctuated dramatically, with regard to
both size and distribution (Milner, 1990). Living in peripheral settings did not provide freedom

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
92 Exposure to infectious pathogens

from disease – some of the highest prevalences of bone infection are in the so-called
“marginal” habitats (Eisenberg, 1986, 1991a, 1991b).
Some evidence suggests that high population density and disease burdens, in combination
with other factors (e.g., warfare), may have contributed to cultural terminations during later
prehistory well before the arrival of Europeans (Eisenberg, 1986; and see Larsen, 1994).
Improved survivorship, coupled with a decline in prevalence of periostitis and osteomyelitis
in Ontario, suggests that populations may have adjusted to high density in this setting
(Katzenberg, 1992). This decline contrasts sharply with other contact era settings where
periosteal infections have been shown to increase in a dramatic fashion (Larsen, Crosby
et al., 2002; Larsen & Harn, 1994; Larsen et al., 2007; Stodder, 1994; Ubelaker, 1994).
Timing of agricultural intensification may explain some of the variation in increasing
prevalences. In contrast to the findings from the analysis of populations from the Eastern
Woodlands, prevalence of periosteal reactions remained unchanged from earlier to later
periods in the Valley of Oaxaca, Mexico (Hodges, 1987, 1989). Unlike most settings in the
Eastern Woodlands, agricultural intensification in the Valley of Oaxaca was accompanied by
neither increased sedentism nor appreciable population growth. Unlike the Eastern Wood-
lands, agricultural development was long and gradual, taking place over several thousand
years. This contrasts with regions of secondary agricultural development, like the Eastern
Woodlands, where maize agriculture was adopted relatively rapidly. Hodges (1987) argues
that the longer period of human – plant interaction may explain some of the differences in
health declines between the Valley of Oaxaca and the Eastern Woodlands.
The shift to agriculture and increased population density in the Nile Valley of Sudanese
Nubia was also not accompanied by an elevation in frequency of bone infectious lesions. In
the X-group intensive agriculturalists (AD 350 – 550) in the Wadi Halfa area, only 12% of
individuals possess nonspecific bone infections, most of which are minor localized periosteal
reactions (Armelagos et al., 1981). This finding is especially surprising because the valley was
densely settled and populations experienced elevated stress (see Chapter 2).
Microscopic examination of femoral cortical bone from X-group individuals indicates a
pattern of fluorescence similar in many respects to tetracycline labeling in modern bone. This
analysis reveals the strong presence of tetracycline – now recognized as a broad spectrum
antibiotic – some 1400 years prior to its medical discovery in the mid-twentieth century
(Bassett et al., 1980; and see Armelagos, 2000; Bassett, 1981; Cook et al., 1989; Keith &
Armelagos, 1988; Maggiano et al., 2003, 2009; Mills, 1992; Nelson et al., 2010). Tetracycline
is highly effective against gram-negative and gram-positive bacteria as well as some other
pathogens; it may have had a highly therapeutic value for ancient Nubians. The source of
tetracycline is unclear, but Bassett and coworkers (1980) suggest that grains – wheat, barley,
and millet – stored in mud bins provided the environmental conditions and nutrients essential
for the growth of Streptomycetes, the bacteria that produce tetracyclines. Alternatively, the
pots used to store grain for brewing ale would have provided the environment suitable for
natural culturing of Streptomycetes (Armelagos, 2000). Similar patterns of bone fluorescence
have been found in skeletal remains in Kulubnarti, Sudan (Hummert & Van Gerven, 1982) and
several settings of the Dakhleh Oasis, Egypt (Cook et al., 1989; Maggiano et al., 2003, 2006,
2009), showing long-term continuance over centuries and significant regional presence. The
application of confocal laser scanning microscopy to the analysis of the Dakhleh Oasis series
in particular provides a record of fluorescence consistent with the presence of tetracycline

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.4 Nonspecific infection and disruption 93

100 mm

Figure 3.8 Confocal laser scanning image of archaeological bone. Osteons are labeled (1)
and (2). Incorporation of tetracycline into the Haversian systems is shown in fluorescent
osteons (2) and lamellae (3). (From Maggiano et al., 2006; reproduced with permission of
authors and Elsevier Ltd.) (A black and white version of this figure will appear in some
formats. For the color version, please refer to the plate section.)

incorporated during life, likely deriving from consumption of stored grains (Figure 3.8).
However, it seems unlikely that therapeutic levels of tetracycline could have been maintained
in life, but bone infection may have been reduced owing to the presence of tetracycline in the
skeleton of affected persons (Maggiano et al., 2003).
Increased population density is not solely dependent upon the adoption of an agricultural-
based economy. A number of regions globally show an increase in sedentism and population
density in the absence of plant or animal domestication. If a chief cause for increasing skeletal
infection is related to demographic factors (i.e., population size and distribution), then
populations undergoing a shift to sedentism in foraging contexts should show similar changes
in infection prevalence as populations adopting agriculture. In order to test this hypothesis,
Lambert and Walker (1991; Lambert 1993, 1994) documented change in prevalence of
periosteal lesions in populations occupying the mainland coast and islands in the Santa
Barbara Channel Islands. Accompanying the shift toward a marine-based economy, there
was an increase in population size and density and decrease in mobility, especially in later
prehistory (Glassow, 1996). By the time of initial contact, native populations in the region had
a level of complexity of social organization and population density that rivaled many
agricultural societies in North America (Arnold, 1992).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
94 Exposure to infectious pathogens

Comparisons of nonspecific periosteal reactions in tibiae reveal a striking increase in


prevalence and severity that peaked during the Late Middle period (AD 580 – 1380) and
declined slightly afterward, in late prehistory. In general, an increase in sedentism and
population size was accompanied by an increase in infection, reflecting a decline in commu-
nity health. In addition to increasing population density, size, and degree of sedentism, other
factors may have contributed to increased infection. Archaeological evidence indicates a clear
pattern of increased exchange between the islands and the mainland, creating the possibility
for the introduction of new infection-causing pathogens. Although diet was rich in protein,
the well-documented increase in interpersonal violence (see Chapter 4) suggests that local
island populations may have become increasingly competitive for limited terrestrial resources
in later prehistory.
The slight decline in bone infection in the late prehistoric period is similar to the pattern
documented for prehistoric Ontario (compare with Katzenberg, 1992), suggesting the possi-
bility of increasing immunities to pathogenic agents associated with high population density.
The Santa Barbara setting shows a continued decline in stature, indicating that health did not
improve. Finally, the region saw an extended period of drought during the Late Middle period
that may have resulted in decreased abundance of food resources, thus contributing to poorer
health and increased infection.
Comparative studies of regional samples of human remains provide an important perspec-
tive on community health in relation to ecological and biocultural variability. Webb (1995)
compared prevalence of infectious lesions in major limb bones from foragers occupying six
regions of Australia: central Murray River valley, Rufus River valley, South Coast, Desert,
northern Tropics, and East Coast. These regions represent highly variable ecological settings,
ranging from tropics to desert. Although temporal comparisons are not available in his study,
the regional perspective represents an important first step toward addressing variability in
nonspecific infection.
Prevalence of nonspecific infection is relatively low throughout Australia, regardless of
region. For example, skeletal samples from coastal areas have a remarkably low prevalence of
infection, the highest being the East Coast right tibiae (6.1%). The highest frequency
of infection in Australia is represented in the Desert group: 16.7% of femora have nonspecific
infections. Reasons for the presence of higher prevalence of infection for this region is not
known. Webb (1995) notes that endemic treponematosis is present in the area, which may
contribute to the higher frequency in the Desert group in comparison with other regions.

3.4.3 Sex and status differences in nonspecific periosteal reactions


There is evidence to suggest that different cohorts of a population are differentially affected
by disease stress. At the late prehistoric Dickson Mounds (Lallo et al., 1978), Moundville
(Powell, 1988), and the Georgia Bight (Larsen, 1998; Larsen & Harn, 1994) sites, prevalence of
nonspecific infections is broadly similar between adult males and females. These similarities
suggest that factors influencing disease prevalence were the same. In the Averbuch and
historic period Georgia Bight populations, males have appreciably higher frequencies of
nonspecific infections than females (Eisenberg, 1986; Larsen, 1998; Larsen & Harn, 1994).
Adult males under the age of 35 years show a tendency to have more unhealed, active lesions
than adult females. Higher prevalence in adult males is also present in other diverse settings:

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.4 Nonspecific infection and disruption 95

Medieval England (Grauer, 1991), intensive agriculturalists from Mesoamerica and Peru
(Danforth et al., 1997; Hodges, 1989; Pechenkina, Vrandenburg et al., 2007), hunter-
gatherer-fishers in coastal southern California (Lambert, 1994), and colonial era African
Americans from New York (Null et al., 2009). Adult males and females show very different
temporal trends in a large series of post-Civil War African Americans in Texas (1869 – 1907;
Davidson et al., 2002). Males show a significant decline, whereas females show an increase.
Overall, males have a considerably higher prevalence than females. Both sexes, however,
display high frequencies of periosteal lesions indicating exposure to conditions conducive to
high levels of infection and infectious disease. This pattern is repeated in other samples of
African Americans, both enslaved and freed in North America (Rathbun & Steckel, 2002; and
see discussion in Blakey, 2001; Barrett & Blakey, 2011).
The tendency for greater male periostitis prevalences in these diverse settings is related to
factors that are unique to specific circumstances. In Spanish Florida, for example, resettlement
of male draft laborers in areas far from home villages may have exposed them to novel
pathogens or other infectious agents (Larsen and Harn, 1994). In the Santa Barbara Channel
Islands region, the propensity for males to participate in highly demanding physical activities
may explain a higher prevalence of periosteal reactions resulting from more blows to their
lower legs than in females (Lambert, 1994).
Several osteological samples show a tendency for adult females to have more nonspecific
lesions than adult males (Martin, 2008; Whittington, 1989). In the prehistoric series from the
La Plata Valley in northwest New Mexico, females show a much higher prevalence of lesions
than males (females¼30.7%; males¼6.2%) (Martin, 2008). At Black Mesa, Arizona, there are
no sex differences in infection prevalence, but females have more severe infections than
males. The finding of greater involvement in females, either with regard to prevalence or
severity, in these Southwestern settings may suggest a greater exposure to pathogens
resulting in infection. In the La Plata Valley, adult females have more skeletal injuries than
males, including cranial depression fractures and other broken elements. In contrast to male
burial, female burial was haphazard and devoid of grave goods. Based on this pattern, Martin
and coworkers (1991, 2001) suggest that adult females lived under suboptimal conditions, at
least in comparison with adult males. In Mórrope, Peru, postcontact females display signifi-
cantly higher periosteal lesion prevalence than their late pre-Hispanic counterparts, suggest-
ing, along with other lines of skeletal and contextual evidence, the relatively greater impact of
conquest on native women in this setting (Klaus & Tam, 2009).
Relatively little paleopathological research is devoted to the important link between status
and health, arguably an important component of complex societies (Powell, 1992a). If elite
members of a society were exempted from activities that would expose them to pathogens, they
should exhibit a relatively lower prevalence of bone infection. Comparisons of high-status with
low-status individuals at Moundville reveals no statistically significant prevalence differences
in infection, suggesting that “status differentiation at Moundville brought no substantial
biological benefits, nor levied any particularly heavy penalties” (Powell, 1992a:88). Compari-
son of skeletal elements of the high-status elites with the two non-elite groups indicates that,
with the exception of the fibula, all long bones from elite individuals have somewhat less
periostitis than those of non-elite individuals. For example, 44.8% of high-status and 51.0% of
low-status tibiae have infections (Powell, 1988: Table 35). Given the vagaries of identification
of high-status individuals as well as potential problems with sample size, these differences may

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
96 Exposure to infectious pathogens

not be meaningful. Other southeastern US Mississippian centers also show no clear differences
between high and low status in infection prevalence (Blakely, 1980, 1988).
In contrast to these late prehistoric settings in the Eastern Woodlands, the picture of social
differentiation and infection prevalence is distinctive at Cahokia, a Mississippian site located
in the American Bottom region of the central Mississippi River valley. The system associated
with the site was the most organizationally complex late prehistoric Mississippian chiefdom in
the Eastern Woodlands (see discussion in Milner, 1990). By the early eleventh century AD, an
elite social stratum had emerged that had differential access to a range of prestige items, and
likely enjoyed a better quality of life than lower classes. At Cahokia Mound 72, analysis of the
remains of 261 individuals from 28 burial features revealed that only 5.3% of high-status
individuals had periostitis, whereas 25.0% of middle-status individuals had lesions (Rose &
Hartnady, 1987).
Comparison of socially elite “shamans” and commoners in the Maitas-Chiribaya culture of
northern Chile (c. AD 1000) reveals that fewer high-status individuals have bone infections
than low-status individuals (prevalence: shamans¼9%; commoner males¼20%; commoner
females¼18%) (Allison, 1984). Although preliminary, these studies indicate that high-status
individuals enjoyed a healthier lifestyle than low-status individuals, at least in these settings.
However, in the Middle Sicán culture, the elite ethnic Sicán appear to have endured signifi-
cantly less infection as well as other forms of stress than the populations that they ruled over
in the Lambayeque Valley Complex (Klaus & Toyne, in press).

3.5 Specific infectious diseases: treponematosis, tuberculosis, and leprosy


3.5.1 Treponematosis
Treponematosis is represented today by four recognized disease syndromes, including ven-
ereal syphilis, nonvenereal (endemic) syphilis (also called bejel), yaws, and pinta (Cook &
Powell, 2012; Hudson, 1965; Mandell et al., 1990; Ortner, 2003; Powell & Cook, 2005a). All
but pinta produce hard tissue responses. Unfortunately, the skeletal lesions of the other
syndromes are so similar that it is virtually impossible to distinguish among them. The
pathogens responsible for the disease are bacterial spirochetes of the genus Treponema,
including T. careteum (pinta), T. pallidum pertenue (yaws), T. pallidum pallidum (venereal
syphilis), and T. pallidum endemicum (endemic syphilis) (Mandell et al., 1990). The physical
appearance of the four species is virtually identical, but molecular (DNA) evidence reveals
differences that warrant subspecies designation (Powell & Cook, 2005a).
Treponemal infection is introduced via the skin or mucous membranes. For venereal
syphilis, the spirochete typically enters the body during sexual contact from lesions on the
genitals. For endemic syphilis and yaws, the pathogen is spread from nongenital lesions on
the arms, legs, or trunk during nonsexual contact between individuals (e.g., physical contact
between children playing). For all three syndromes, the infection spreads throughout the body
via the circulatory system. Congenital transmission of venereal syphilis or yaws involving
passage of the spirochete from the mother transplacentally to the fetus is well documented
(Hudson, 1965; Ortner, 2003). In living populations, pinta, endemic syphilis, and yaws have
been found to be especially prevalent in rural settings with poor sanitation. Additionally, in
temperate to hot climates, individuals typically wear relatively little clothing, which facilitates

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.5 Specific infectious diseases: treponematosis, tuberculosis, and leprosy 97

the spread of infection through direct contact with infected skin abrasions and cuts. Venereal
syphilis characteristically appears in populations with higher levels of sanitation, such as in
urban settings in Western countries. These groups tend to be relatively more fully clothed,
which provides fewer opportunities for the spread of infection via skin contact in the manner
typical of nonvenereal syphilis, pinta, and yaws.
The skeletal manifestations of the treponematoses are described in the paleopathological
literature in considerable detail (Hackett, 1976; Ortner, 2003; Powell, 1988; Powell & Cook,
2005b). Yaws is most commonly represented by the inflammatory response of the periosteum
surrounding the bones of the forearm, hand, and of the lower limb bones. In the most severe
form, repeated episodes of periosteal reaction and remodeling may result in hypertrophy of
the anterior crests of the tibia and an appearance of bowing called “saber-shin.” Bone surfaces
with close proximity to skin – such as the cranial vault and the anterior tibia – also may
express active lesions or pitted defects from gummatous granulomas. Destructive nasal and
hard palate changes may occur, but are less prevalent in yaws than in venereal or endemic
syphilis.
Skeletal changes associated with endemic syphilis are similar to those that develop in yaws.
In the tertiary stage of the disease, periostitis and gummatous granulomas may develop in the
cranial flat bones and tibiae, and the tibiae take on a saber-shin appearance (Figure 3.9).
Destructive lesions of the face, especially in the nasal region, may also develop. Venereal
syphilis results in virtually the same bone lesions as endemic syphilis, including extensive
cranial vault lesions and periosteal inflammation of lower limb long bones. Other destructive
changes may be present in elbow, hip, and knee joints.
Tertiary-stage venereal syphilis may involve abundant osteosclerotic skeletal responses,
characterized by gummatous destruction of bone. The skull is frequently affected, especially
the nasal area and flat vault bones (Figure 3.10). The frontal bone typically expresses healed
star-shaped, gummatous lesions called caries sicca. Over the course of time, these lesions may
coalesce to form an ectocranial surface with a high degree of irregular topography.
Congenital syphilis results in distinctive skeletal changes, such as osteochondritis (poor
bone formation in areas of endochondral ossification), periostitis, and osteomyelitis. Dental
changes do not occur in the nonvenereal treponematoses because the teeth are fully or nearly
fully formed in the secondary and tertiary stages. In up to 30% of congenital syphilitic
children, pathognomonic modifications may occur in the forming permanent anterior teeth,
specifically the characteristic malformation of incisors in which the crown is unnaturally
constricted at the occlusal margin (called Hutchinson’s incisors, or screwdriver teeth), sharp
hypoplastic defects on canines, and the anomalous patterning of first molars involving
abnormally closely spaced cusps (Moon’s molars) or the presence of a plane-form hypoplasia
cutting into the bases of all cusps (Fournier molars) (Hillson et al., 1998) (Figure 3.11). These
lesions are extremely rare in New World archaeological remains dated before 1492, and are
more common in historic populations with documented evidence of venereal syphilis (Cook,
1994; Jacobi et al., 1992; Lambert, 2006; Mansilla & Pijoan, 1995; Nystrom, 2011; Ortner,
2003). Several possible cases from Old World contexts have been identified (Erdal, 2006;
Henneberg & Henneberg, 1994; Palfi et al., 1994), indicating that venereal syphilis in Europe
may predate the post-Columbian era.
Presence of treponematosis with special attention to venereal syphilis in past populations,
especially in the New World, has been debated for well over a century. Analysis of human

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
98 Exposure to infectious pathogens

Figure 3.9 Periosteal inflammation of tibiae (treponematosis); Irene Mound, Georgia. The
tibia in the middle is nonpathological. (From Powell, 1990; reproduced with permission of
author and American Museum of Natural History.)

remains from Tennessee and Kentucky by the American Civil War surgeon Joseph Jones
revealed “the unmistakable marks of the ravages of syphilis” (1876:66). The presence of
“diseased, enlarged, and thickened” long bones convinced him that syphilis was widespread
in the region, that it had an exclusively New World origin, and that it must have been
imported to Europe via the West Indies (Jones, 1876:67). This discussion continues a debate
that is still ongoing; namely, the origin of venereal syphilis, New World or Old World (Baker &
Armelagos, 1988; Cook & Powell, 2012; Dutour et al., 1994; Harper et al., 2011; Merbs, 1992;
Powell & Cook, 2005a, 2005b). Owing to the overlapping symptoms and skeletal manifest-
ations of these three treponemal diseases and the far more extensive record of research on
North American skeletal remains than in Europe or other areas of the Old World, the issue
remains contentious.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.5 Specific infectious diseases: treponematosis, tuberculosis, and leprosy 99

Figure 3.10 Treponematosis (caries sicca) on cranial vault from Early Woodland (900 BC–
AD 200), Wilhoite site, Tennessee. (Photograph by Tracy K. Betsinger.)

One or more syndromes of treponematosis were present in both the Old World and the New
World well prior to AD 1492 (Baker & Armelagos, 1988; Cook & Powell, 2012; various in
Powell & Cook, 2005b; and many others). Cases of treponematosis are sparse in the Old World,
but a growing number of cases, some aided with histological evidence, are expanding that
record for Europe and East Asia (Brothwell, 2005; Cole & Waldron, 2011; Dutour et al., 1994;
Erdal, 2006; Henneberg & Henneberg, 1994; Henneberg et al., 1992; Mays et al., 2003;
Mitchell, 2003; Roberts, 1993; Roberts & Manchester, 2005; Rothschild & Rothschild, 1995;
Suzuki, 1982 – 1984; von Hunnius et al., 2006). Treponematosis is widespread in Australia
(Hackett, 1936, 1976; Webb, 1995) and in some areas of the Pacific (e.g., Marianas: Hanson,
1988; Stewart & Spoehr, 1952; Stodder et al., 1992). While these cases show the Old World
presence of some form of treponematosis, owing to the uncertainly of either clear diagnosis or
good chronological control of cases from archaeological settings, evidence supporting an Old
World origin for venereal syphilis remains elusive (Harper et al., 2011).
The evidence for treponematosis from the New World is quite abundant, having been
identified with regard to both a range of case studies and population differential diagnoses,
especially in North American contexts. In addition to earlier studies (see reviews in Baker &
Armelagos, 1988; Powell & Cook, 2005a), examples of treponematosis are associated
with highly diverse settings (Arriaza, 1995; Bruwelheide et al., 2010; Cybulski, 1980, 1990;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
100 Exposure to infectious pathogens

Figure 3.11 Development deformity of anterior mandibular dentition (enamel hypoplasia


and Hutchinson’s incisors) associated with congenital syphilis; Eaton Ferry Cemetery, North
Carolina. (From Lambert, 2006; reproduced with permission of author and Fundação
Oswaldo Cruz.) (A black and white version of this figure will appear in some formats. For
the color version, please refer to the plate section.)

Lahr & Bowman, 1992; Lambert, 1994; Marden & Ortner, 2011; Owsley & Rose, 1997;
Pechenkina, Vrandenburg et al., 2007; Powell & Cook, 2005a, 2012; Sandford et al., 2002;
Schermer et al., 1994; Stodder, 1994; Stodder & Martin, 1992; Walker & Lambert, 1989;
Wright, 2006; various in Powell & Cook, 2005b). In many contexts, the skeletal manifestation
appears intermediate between those of the two modern endemic syndromes. Considering the
evolutionary nature of human infectious disease over centuries and across host populations of
differing genetic composition, these departures from the modern pattern are not unexpected.
The most extensive record of New World treponematosis data is from precontact Eastern
North America, especially in later prehistoric settings involving sedentary communities
practicing farming. As discussed earlier (see Chapter 2), it is these settings that provide ideal
conditions for the maintenance and spread of disease-causing pathogens. In the lower Illinois
River valley, a pattern of proliferative bone infection that is strongly suggestive of endemic
treponematosis rather than venereal syphilis is present (Cook, 1976, 2007). There is a high
frequency of tibial periostitis affecting adult males and females alike, progressively increasing
with age. The prevalence is especially elevated in later prehistoric, maize-dependent groups
with high population density. Although variable across populations, the pattern has been
described in prehistoric (mostly late) human remains from other groups in a wide range of
settings, but especially in the American Midwest and Southeast (Bogdan & Weaver, 1992;
Bullen, 1972, 1973; Cassidy, 1984; Cook, 2005; Danforth et al., 2007; Dickel, 1991; Eisenberg,
1986, 1991a, 1991b; Garner, 1991; Hutchinson, 1993, 2004; Hutchinson & Mitchem, 1996;
Hutchinson & Norr, 2006; Hutchinson & Richman, 2006; Hutchinson & Weaver,
1998; Hutchinson et al., 2005, 2007; Işcan & Miller-Shaivitz, 1985; Lambert, 2000b; Lewis,
1994; Miller-Shaivitz & Işcan, 1991; Milner, 1983, 1992; Milner & Smith, 1990; Monahan,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.5 Specific infectious diseases: treponematosis, tuberculosis, and leprosy 101

Figure 3.12 Stellate lesions on adult frontal bone (treponematosis); Tierra Verde, Florida. (From
Hutchinson, 1993; reproduced with permission of author and John Wiley & Sons, Inc.)

1995; Powell, 1986, 1988, 1989, 1990, 1991a, 1991b, 1992a, 1992b, 1994; Powell et al., 2005;
Reichs, 1989; Robbins, 1978; Ross-Stallings, 1989; Smith, 2008; Smith et al., 2011; Weaver
et al., 2005; Wilson, 2005).
A number of these settings provide details on lesion morphology and other characteristics that
strongly suggest the New World record reflects multiple forms of endemic treponematosis (Powell
& Cook, 2005a). Hutchinson (1993) found proliferative periosteal apposition on long bones
(especially tibiae) and stellate lesions in crania from late prehistoric and early contact contexts
in the central Florida Gulf coast. For example, in the sixteenth century component of the Tatham
Mound sample, three crania exhibited healed stellate lesions (Figure 3.12). Similar lesions are
present in crania from the postcontact Weeki Wachee Mound and precontact components of the
Safety Harbor and Tierra Verde sites. Although saber-shin tibiae are also present in these samples,
stellate scars appear to be the best single criterion for documenting the presence of endemic
treponematosis (Hutchinson, 1993; and see Hutchinson & Norr, 2006; Hutchinson & Richman,
2006; Hutchinson et al., 2007; Milner & Smith, 1990; Powell, 1990, 1991a) (Figures 3.10, 3.12).
How life-threatening or debilitating treponemal disease was in these earlier societies is
unknown. Owing to the apparent endemic nature of the disease, it may not have been a
primary cause of mortality. Based on the high degree of healing, Powell (1988) argues that
populations at Moundville had successfully adapted to the disease. While this may be the case,
nevertheless, the presence of the characteristic lesions indicates that the disease imposed a
significant health burden and resulted in no small amount of discomfort for those affected. In
his description of native populations in North Carolina in the early eighteenth century, the
explorer John Lawson noted that natives in the eastern part of the colony “. . .have a sort of
Rheumatism or Burning of the Limbs, which tortures them grievously, at which time their Legs
are so hot, that they employ the young People continually pour water down them. . .This not
seldom bereaves them of their Nose. I have seen three or four of them render’d most miserable
Spectacles by this Distemper. Yet, when they have been so negligent, as to let it run on so far

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
102 Exposure to infectious pathogens

without curbing of it; at last, they make shift to patch themselves up, and live for many years
after. . .” (Lawson, 1967:231). These descriptions correspond well with modern clinical
descriptions of deep leg pain and orofacial lesions produced by treponemal disease (Hackett,
1951; Hudson, 1958).
Some examples of nonspecific periosteal reaction observed in archaeological skeletons
likely have a treponemal origin, especially in individuals who also display the classic path-
ognomonic lesions (e.g., stellate scars on crania and/or highly proliferative bone on tibiae
diaphyses). In cases where distinctive symptoms are absent, the delineation between non-
specific infection (periostitis) and endemic treponematosis is difficult to determine.
The well-documented paleopathological diagnoses of treponematosis in North American
archaeological settings make it clear that the disease was well established in the New World
prior to the arrival of Europeans. In the lower American Midwest and Southeast, regions that
are subtropical or that experience high seasonal humidity and temperatures, there appears to
be a cline of skeletal expression from the hotter, more humid coastal regions to the relatively
drier interior regions, resembling modern inter-regional treponemal clines spanning “classic”
yaws and “classic” endemic syphilis in central and southern Africa (Basset et al., 1994;
Froment, 1994; Grin, 1956).
Identification of Treponema DNA in archaeological skeletons is highly problematic.
T. pallidum pallidum has been successfully identified in only one setting (Kolman et al.,
1999; but see Barnes & Thomas, 2006), indicating that molecular approaches to understanding
this disease are extremely sketchy, at best (Bouwman & Brown, 2005; Buikstra, 2010; von
Hunnius et al., 2007; Wilbur & Stone, 2012). At least with respect to understanding the history
of this disease, it is, for the time being, largely dependent upon the skeletal record.
Variation by sex or status in skeletal populations is difficult to assess, especially given the
vagaries of diagnosis. Where probable cases of treponematosis have been identified, the
prevalences in adult males and females are broadly similar (Cook, 1976; Powell, 1988; Powell &
Cook, 2005a; Smith et al., 2011; but see Powell, 1990). Powell’s (1988, 1992b) investigation of
the Moundville skeletal series indicates no clear distinction between status groups. In contrast,
comparison of elite and non-elite individuals from late prehistoric (Mississippian period)
stratified societies in eastern Tennessee reveals high-status individuals as having significantly
fewer skeletal lesions associated with treponematosis than low-status individuals (Smith et al.,
2011). This clear dichotomy in disease prevalence suggests that higher-status individuals may
have enjoyed circumstances conducive to relatively better health than low-status individuals.
Alternatively, more robust health may have facilitated social advancement. Current models of
disease epidemiology suggest, however, that the former explanation is more likely.

3.5.2 Mycobacterial infection: tuberculosis


Tuberculosis is among the oldest infectious diseases in Homo sapiens. Genetic analysis
suggests that the human form of the pathogen, Mycobacterium tuberculosis, may be several
million years old, originating well before the emergence of Mycobacterium bovis (Buikstra,
2010; Stone et al., 2009). It has long been assumed that the human strain of M. tuberculosis
evolved from non-human strains (e.g., M. bovis). However, new analyses show that the more
likely evolutionary sequence is the human strain giving rise to non-human strains, perhaps
during the period of domestication (Brosch et al., 2002). This is important to know because it

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.5 Specific infectious diseases: treponematosis, tuberculosis, and leprosy 103

provides a long-term context of disease history, human – animal interaction, and co-
evolution of pathogens.
Tuberculosis today is comprehensively global, infecting nearly one-third of the population,
second only to HIV as a cause of death in human populations. In 2005 alone, there were
nearly nine million new cases with most in Asia and Africa, making it a leading cause of death
(World Health Organization, 2007). Unlike treponematosis, which produces proliferative bone
apposition, tuberculosis progressively destroys bone tissue, and is commonly expressed as
erosive vertebral lesions of the lower back (lower thoracic and lumbar vertebrae) and resorp-
tive and slight proliferative changes of the pleural (internal) surfaces of ribs (Ortner, 2003;
Palfi et al., 1999; Roberts & Buikstra, 2003; Roberts et al., 1994). Tuberculosis is caused by
organisms from the Mycobacterium tuberculosis complex, here simply called “tuberculosis.”
The most common cause of tuberculosis today in humans is the bacteria M. tuberculosis and
M. bovis. The primary mode of transmission for the former is by breathing airborne microbes,
usually in droplets introduced by sneezing or coughing, and the primary mode for the latter is
via consumption of milk and meat from infected animals or by swallowing infected sputum of
the intestinal tract from animal products infected by the bacterium (Gordon & Mwandumba,
2008; Roberts, 2011).
The infection pathway is usually through the respiratory tract, resulting in a primary
infection in lung tissue and subsequent secondary infection in regional hilar lymph nodes
(Ortner, 2003). Over a period of years, the bacilli may spread to skeletal tissues via the
circulatory system, with a propensity for iron-rich hematopoietic marrow and cancellous
bone. The vertebrae, ribs, sternum, and (for subadults) long bone metaphyses are especially
favored sites of secondary infection because of the presence of a rich blood supply and the
scarcity of phagocytic cells (Hopewell, 1994; Ortner, 2003; Thijn & Steensma, 1990). The
process can result in extensive destruction of cancellous bone, most commonly in vertebral
bodies (Figure 3.13). With the loss of bone mass, vertebral bodies may collapse and fuse. The
resulting severe kyphosis is called “Pott’s disease” after the original description by Sir
Percivall Pott (1779) (Figure 3.14). However, any bone or joint can be involved in tubercular
infections (Berney et al., 1972; Ortner, 2003).
The proliferative lesions on the visceral surfaces of ribs are presumed by some to be
pathognomonic for tuberculosis, but the cause remains unclear and may reflect secondary
infection downstream from the original tubercular infection or some other infection
altogether (Kelley & Micozzi, 1984; Pfeiffer, 1991; Roberts et al., 1994, 1998; Wilbur et al.,
2008) (Figure 3.15). Examination of skeletal remains from individuals of known cause of
death shows individuals with rib lesions as also having tuberculosis (Santos & Roberts, 2001).
Individuals with non-tubercular pulmonary disease sometimes have similar rib lesions.
Therefore, all rib lesions of this type should not necessarily be interpreted as diagnostic of
tuberculosis (Roberts, 1999; Roberts et al., 1994), and may reflect other respiratory infections,
such as some form of pneumonia (Lambert, 2002b). Nevertheless, rib lesions record an
important possible signal of tuberculosis infection in past populations but the pathology data
should be interpreted cautiously (Pfeiffer, 1991; Raff et al., 2006; Roberts et al., 1998; Santos &
Roberts, 2001; and see Matos and Santos, 2006) and perhaps confirmed via analysis of aDNA
(Müller et al., 2014).
In the Old World, skeletal evidence for tuberculosis has been traced to six to seven thousand
years ago in Europe (Nicklisch et al., 2012), but with significant levels of prevalence occurring

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
104 Exposure to infectious pathogens

Figure 3.13 Destructive lesions on thoracic vertebral bodies pathognomonic of tuberculosis;


Lambayeque, Peru. (Klaus et al., 2010; reproduced with permission of author and
Elsevier Ltd.) (A black and white version of this figure will appear in some formats. For the
color version, please refer to the plate section.)

considerably later, especially in the later Medieval period when populations began to be
concentrated in urban or otherwise settled areas (Canci et al., 1996; Gladykowska-Rzeczycka,
1999; Roberts and Buikstra, 2003; Stone et al., 2009). In many areas of the Old World, there is
limited or no record of tuberculosis. The paucity of data may simply reflect the fact that
bioarchaeological research is more comprehensive in Europe than elsewhere.
Most early authorities argued that the disease was absent in the New World prior to
European contact (Hrdlička, 1909). However, the growing skeletal record in the New World
reveals an abundance of pathology indicating that this illness appeared by the first millen-
nium AD (Drake & Oxenham, 2012; Klaus, Wilbur et al., 2010; Roberts, 2012; Roberts &
Buikstra, 2003; Stone et al., 2009). Moreover, the identification of acid-fast bacilli and soft-
tissue tubercular lesions (Allison et al., 1973, 1981), and especially M. tuberculosis DNA
in precontact Peruvian and Chilean mummies (Arriaza et al., 1995; Konomi et al., 2002;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.5 Specific infectious diseases: treponematosis, tuberculosis, and leprosy 105

Figure 3.14 Destructive lesions on bodies of thoracic and lumbar vertebrae, resulting in
kyphosis of the spine (Pott’s disease, tuberculosis); Norris Farms, Illinois. (Photograph by
George R. Milner and the Illinois State Museum.) (A black and white version of this figure
will appear in some formats. For the color version, please refer to the plate section.)

Figure 3.15 Tuberculosis rib lesions; Terry Collection. (Photograph by Charlotte A. Roberts.)
(A black and white version of this figure will appear in some formats. For the color version,
please refer to the plate section.)

Salo et al., 1994) and skeletons (Bos et al., 2014; Braun et al., 1998; Kaestle, 2010; Klaus,
Wilbur et al., 2010; Raff, 2008) demonstrates its presence in the New World in the pre-
Columbian era. The significant Old World distribution of the disease via detection of M.
tuberculosis DNA is also indicated in archaeological remains from a range of contexts (Baron
et al., 1996; Boros-Major et al., 2011; Crubézy et al., 2006; Fletcher et al., 2003; Haas et al.,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
106 Exposure to infectious pathogens

2000; Hershkovitz et al., 2008; Matheson et al., 2009; Mays et al., 2001; Murphy et al., 2009;
Nicklisch et al., 2012; Nuorala, 1999; Roberts, 2012; Spigelman & Lemma, 1993; Suzuki,
2013; Szelekovszky & Marcsik, 2010; Taylor et al., 1996, 2005; Zink et al., 2005; various in
Pálfi et al., 1999). The earliest purported molecular (DNA) evidence is from Neolithic settings
dating to c. 9000 years ago from Israel (Hershkovitz et al., 2008; but see Wilbur et al., 2009)
and c. 7000 years ago in Germany (Nicklisch et al., 2012).
Destructive vertebral or proliferative rib lesions are identified in archaeological skeletal
series representing a diversity of groups worldwide, including the Middle East (Baker, 1997;
Buikstra et al., 1993; Morse, 1967; Ortner, 1979; Ortner & Frohlich, 2008; Roberts & Buikstra,
2007; Strouhal, 1991; Zias, 1991), northern and central Europe (Arcini, 1999; Bennike, 1985,
1999; Blondiaux et al., 1999; Evinger et al., 2011; Formicola et al., 1987; Inglemark, 1939;
Jankauskas, 1999; Manchester, 1991; Mays & Taylor, 2003; Mays et al., 2001; Pálfi &
Marcsik, 1999; Roberts, 2011; Roberts & Buikstra, 2003; Waldron, 1993), Greece (Angel,
1984), Asia (Murphy et al., 2009; Oxenham et al., 2005; Pechenkina, Benfer et al., 2007;
Suzuki, 1991, 2013; Suzuki & Inoue, 2007; Suzuki et al., 2008; Tayles & Buckley, 2004), and
elsewhere (Ortner, 2003; Stone et al., 2009). In the New World, a spate of reports document the
presence of a disease strongly resembling tuberculosis in South America (Allison, 1984;
Allison et al., 1981; Arriaza et al., 1995; Buikstra & Williams, 1991; Klaus, Wilbur et al.,
2010; Wilbur et al., 2008). In North America, instances of tuberculosis are restricted to
primarily eastern settings (Buikstra, 1977b; Buikstra & Cook, 1978, 1981; Clabeaux, 1977;
Cook, 1984; Danforth et al., 2007; Eisenberg, 1986, 1991a, 1991b; Hartney, 1981; Katzenberg,
1977; Lambert, 2000b; Milner, 1983, 1992; Milner & Smith, 1990; Murray, 1989; Pfeiffer,
1984, 1991; Pfeiffer & Fairgrieve, 1994; Powell, 1988, 1990, 1991a, 1991b, 1992a, 1992b,
2000; Rathbun et al., 1980; Saunders et al., 1992; Widmer & Perzigian, 1981) and the
Southwest (Bruwelheide et al., 2010; El-Najjar, 1979; Fink, 1985; Hooton, 1930; Micozzi &
Kelley, 1985; Stodder, 1994; Stodder & Martin, 1992; Stodder et al., 2002; Sumner, 1985). In
addition, other cases have been documented on the Northwest Coast (Cybulski, 1978, 1990)
and Great Plains (Mann et al., 1994; Owsley & Rose, 1997; Palkovich, 1981; Williams, 1994;
Williams & Snortland-Coles, 1986). However, in Mesoamerica, a very notable lack of tuber-
culosis is well recognized (Wilbur et al., 2008).
Most paleopathological studies report on isolated cases of tuberculosis or tuberculosis-
like infections. The detailed study of prevalence, pattern, and lesion morphology in a
limited number of series provides important details on the history of the disease. In a
comprehensive study of a temporal sequence of human remains from the lower Illinois
River valley in west-central Illinois, biocultural reconstruction and interpretation of change
in pattern and prevalence of resorptive lesions indicates the presence of tuberculosis
(Buikstra and Cook, 1981; also Buikstra, 1977b; Buikstra and Cook, 1978; Cook, 1984,
2007). Over the course of the sequence, beginning in the Middle Woodland (150 BC – AD
400), followed by the Late Woodland (AD 400 – 1050) and Mississippian (AD 1050 – 1400)
periods, the region saw an increase in population density and sedentism, especially when
maize agriculture became well established during the eleventh century AD. No clear
evidence of tubercular resorptive lesions is present in the Middle or Late Woodland periods.
In the Mississippian period, however, there is a clear reorientation in disease pattern.
Human remains from the Mississippian-period Yokem and Schild sites possess destructive
vertebral body lesions consistent with an etiology of tuberculosis. The Mississippian-period

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.5 Specific infectious diseases: treponematosis, tuberculosis, and leprosy 107

adults show a high young adult mortality and an equitable distribution among adult males
and females.
At Moundville, circumstances for the introduction and spread of tuberculosis are similar to
west-central Illinois (Powell, 1988, 1991a, 1991b, 2000). In comparison with Illinois, vertebral
lesions are rare: only three adults are affected in this manner, and only one adult displays
classic vertebral body destruction. No crania, hips, or knees possess resorptive lesions con-
sistent with tuberculosis. However, the presence of a large number of pleural rib lesions may
indicate a broader presence of the disease in this population. If so, this means that rib lesions
may be an early manifestation of tuberculosis if it first enters the lungs.
At the Irene Mound site, three individuals show osteolytic vertebral lesions. One individual
has extensive destruction of the sacroiliac joint without remodeling, two individuals have
periostitis on the anterior scapular bodies, and one individual has periostitis on the pleural
aspect of the sternum (Powell, 1990, 1991b). Additionally, eight of 10 Irene Mound individ-
uals with other tubercular lesions have subtle periosteal apposition on the pleural aspects of
the ribs. Although the skeletal changes at the Irene Mound site are not as profound as those
observed in the American Midwest (compare with Buikstra & Cook, 1981; Milner & Smith,
1990), they nevertheless fit the expected profile of tuberculosis.
Among the most comprehensive yet arguably imperfect bioarchaeological records of
tuberculosis is in the precontact Andean region, with some of the earliest cases in the
New World around 2000 years ago in the Pisco Valley (Peru) and Atacama region (Chile)
(Allison et al., 1981). These and other contexts from the region provide a number of sites
with tuberculosis, but most sites reporting disease have only one or two cases. Indeed, the
largest is represented by 10 individuals from Estuqueña (Peru) (Buikstra & Williams, 1991).
Therefore, unlike Europe where clear diachronic trends show an increase in the Medieval
period relative to earlier periods, the diachronic record in the Andean region needs to be
developed further (Klaus, Wilbur et al., 2010). Nonetheless, the skeletal record suggests that
tuberculosis peaked around AD 1000. This appears to have coincided with the appearance of
complex societies – chiefdoms and states – and associated population size increase, popula-
tion density, economic intensification, and dietary changes, promoting the infection (Klaus,
Wilbur et al., 2010). These regional analyses provide compelling evidence that, as in living
populations, the synergy of living circumstance – especially population aggregation, com-
promised health, and changed dietary foci – served to promote conditions conducive for the
pathogen to thrive in the past. This pattern is repeated worldwide where data for disease
prevalence and abundant archaeological context are available (Roberts & Buikstra, 2003;
Roberts & Cox, 2003, 2007).
Importantly, the strain or strains of M. tuberculosis in the precontact New World are
American-specific, and distinctive from the strain documented in native populations in
historic times. This finding is consistent with the hypothesis that the American-specific strains
of the tubercular pathogenic organisms were replaced by a European strain after contact (Bos
et al., 2014). One of the more provocative possibilities that Klaus, Wilbur, and coworkers
(2010) consider from their lack of identification of M. tuberculosis aDNA in a wide variety of
skeletons could indicate the infectious agent in the western hemisphere before European
contact was not tuberculosis at all, but rather, a closely related form, such as M. kansasi,
that can produce identical lesions in the human skeleton. It could also mean that the aDNA
simply is not preserved. These possibilities require further scrutiny, especially regarding the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
108 Exposure to infectious pathogens

designs of assays employed to identify tuberculosis versus other very closely related
organisms.
Distribution of tuberculosis by sex in these groups shows considerable variation (Buikstra &
Cook, 1981; Powell, 1988, 1992a). In some settings, there appears to be a male bias. In the
Andean setting, for example, the higher male prevalence may be related to exposure relating
to behaviors or dietary variation occurring prior to adulthood (Klaus, Wilbur et al., 2010). In
the Irene Mound population, however, female prevalence is twice that for males (Powell,
1991a), but this could be an artifact of the composition of the sample (Larsen, 1984, for
discussion of age bias). In terms of status differences, the only systematic analysis indicates no
apparent relationship between tuberculosis and rank in at least one setting in the American
Southeast (Powell, 1988).

3.5.3 Mycobacterial infection: leprosy


Coexisting with tuberculosis in many regions of the world today is leprosy. Skeletal lesions of
both diseases are readily identifiable in skeletal series, and DNA from both diseases extends
this coexistence into prehistory (Donoghue et al., 2010). Like tuberculosis, leprosy (also called
Hansen’s disease) is a chronic infection caused by the acid-fast, gram-positive bacillus,
Mycobacterium leprae (Boldsen, 2008; Carmichael, 1993; Lynnerup & Boldsen, 2012; Ortner,
2003). The bacilli are transmitted by inhalation (Davey, 1974; Gelber, 2008; Roberts, 2011).
Unlike tuberculosis, leprosy is not readily communicable – only about 5% of persons who are
exposed contract the disease. Most who acquire the infection have been in prolonged contact
with infected individuals. The incubation period is very long, occurring over the course of
years or even decades (Steinbock, 1976). The disease is usually not fatal, but neurological
complications are long-lasting.
A variety of manifestations of the disease are possible (Scollard et al., 2006). As discussed
later, in advanced stages, it is accompanied by severe disfigurement of the skin, soft tissues,
and skeleton. Unlike tuberculosis, the disease is not present in significant numbers globally.
Only 220 000 cases were reported on the world registry for the disease, and it is estimated that
the disease is declining dramatically (Stone et al., 2009). Today, the disease is limited mostly to
tropical and subtropical regions of Africa, Asia, and South America, but in the past it was
probably much more widespread, extending as far north as the Arctic Circle (Carmichael,
1993).
After initial infection, M. leprae multiplies slowly, usually in sheaths of peripheral nerves.
The primary stage of the disease involves a loss of sensation due to inadequate innervation.
Thus, minor damage to the skin – as in a cut or scrape – does not elicit a pain response. Owing
to the poor blood supply and repeated injury of affected tissues, healing is hampered, resulting
in localized infection. Over time, various parts of the body, especially toes, fingers, and nasal
tissue, are disfigured or lost entirely. The disease usually affects the skeleton in advanced
cases (see later). The degree of infection and involvement of the body and skeleton is also
influenced by the immune system. If a person is immunosuppressed, the degree of infection
will be elevated.
Its presence in the Old World is confirmed by identification of M. leprae DNA in osteo-
logical tissues (Donoghue et al., 2002, 2010; Haas et al., 2002; Matheson et al., 2009; Rafi
et al., 1994; Spigelman & Donoghue, 2002; Spigelman & Lemma, 1993; Taylor et al., 2000,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.5 Specific infectious diseases: treponematosis, tuberculosis, and leprosy 109

2013; and others). Unlike treponematosis and tuberculosis, leprosy was not present in the New
World prior to European contact. In all likelihood, the disease was introduced to the New
World during the early colonial era (Ortner, 2003). However, no diagnoses in New World
settings have been undertaken.
Beginning largely with the work of the Danish physician and paleopathologist Vilhelm
Møller-Christensen (1961, 1978) on human remains recovered from Medieval leprosy hospital
cemeteries in Denmark (and see Andersen, 1969; Bennike, 2002; Linderholm & Kjellström,
2011), the skeletal manifestations of leprosy are well delineated. His pioneering studies
contributed to the modern diagnosis of the disease in living populations and serve as a
baseline for understanding the disease in archaeological contexts.
Excavations at St. Jorgensgard, a Danish church cemetery (c. AD 1250 – 1550) near
Næstved, resulted in the recovery of about 650 skeletons (Møller-Christensen, 1961, 1978).
This is a unique series for bioarchaeological investigation because all individuals represent
those who had been admitted into the leprosy hospital in order to isolate them from other
members of the population at large (Møller-Christensen, 1978). Møller-Christensen’s exten-
sive studies of this series revealed a distinctive “facies leprosa” or rhinomaxillary syndrome
involving atrophy of the nasal and maxillary regions, alveolar resorption, and anterior tooth
loss (Figure 3.16a). Additionally, hand and foot elements are atrophied and shortened
(Andersen et al., 1992) (Figure 3.16b). Considerable evidence from Denmark, England, and
other European settings shows that the disease first appears in the second millennium BC, has
a significant presence in the early Medieval period, and increases to its maximum in the late

(a)

Figure 3.16a Rhinomaxillary syndrome in leprosy; Early Medieval, Lauchheim, Germany.


(From Boldsen, 2008; reproduced with permission of author and John Wiley & Sons, Inc.)
(A black and white version of this figure will appear in some formats. For the color version,
please refer to the plate section.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
110 Exposure to infectious pathogens

(b)

Figure 3.16b Metatarsal atrophy of the feet (leprosy); Late Medieval/Early Post-Medieval,
Armoy, Ireland. (From Murphy, 2007; reproduced with permission of the author.)

Medieval period (Antunes-Ferreira et al., 2014; Arcini, 1999; Belcastro et al., 2005; Blondiaux
et al., 1994, 2002; Boldsen, 2005, 2008; Boldsen & Mollerup, 2006; Brothwell et al., 2000;
Haas et al., 2002; Judd & Roberts, 1998; Kjellström, 2012; Manchester, 1981; Mariotti et al.,
2005; Murphy, 2007; Murphy & Manchester, 2002; Pálfi, 1991; Pálfi et al., 2002; Roberts,
2002, 2011; Roberts & Cox, 2003; Rubini & Zaio, 2009; Strouhal et al., 2002). While not as
widespread as in Europe, other leprotic skeletons have been described from the Nile valley
region (Dzierzykray-Rogalski, 1980; Møller-Christensen & Hughes, 1966; Molto, 2002; Smith,
1908), Israel (Manchester, 1993; Zias, 2002; Elliot ), Turkey (Angel, 1969), Uzbekistan (Blau &
Yagodin, 2005), and India (Robbins Schug et al., 2009). The remains from Turkey date to
2700 – 2300 BC and those from India date to 2500 – 2000 BC, representing the oldest probable
cases of leprosy. As with tuberculosis, the presence of leprosy is associated with settings
involving population crowding and poor living circumstances generally. In northern Europe,
the cases are drawn largely from small villages in rural settings, similar to what is seen today.
Temporal documentation in Europe, where the data are the richest, suggests that the disease
declines after c. AD 1300 and largely disappears by the nineteenth century, perhaps due to the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.6 Specific infectious diseases: vectored infections 111

rise of tuberculosis and circumstances relating to cross-immunity of the two diseases (Dono-
ghue et al., 2010; Leitman et al., 1993; Manchester, 1991; Roberts & Buikstra, 2003).
Other skeletal pathology found in leprous individuals includes cribra orbitalia, periostitis on
tibiae and fibulae, and maxillary sinusitis (Andersen, 1969; Boocock et al., 1995; Møller-
Christensen, 1978; Roberts et al., 2002), although these conditions alone are not diagnostic of
leprosy. Increased prevalence of maxillary sinusitis, for example, appears to accompany
elevation in air pollution and the confines of urban living in later Medieval England (Lewis
et al., 1995) and other settings (Roberts, 2007). Some individuals display dental changes
whereby crown bases of maxillary incisors are concentrically constricted (Roberts, 1986). The
presence of malformed teeth indicates that the infection occurred early in childhood.
Additional examples of leprosy from archaeological contexts are from mostly isolated
skeletons from other Old World localities (Boocock et al., 1995; Hirata et al., 2000; Lewis
et al., 1995; Magilton et al., 2008; Manchester, 1991; Manchester & Roberts, 1989; Møller-
Christensen & Inkster, 1965; Ortner, 2003; Tayles & Buckley, 2004; Trembly, 1995, 2002; Zias,
1991), but the bioarchaeological evidence is far less profuse in these regions than in Denmark
and England. Manchester and Roberts (Manchester, 1991; Manchester & Roberts, 1989; and
see Roberts et al., 2002) assessed skeletal, archaeological, and archival evidence of leprosy in
Britain and conclude that, as in Scandinavia, the disease was endemic. They argue that
following the introduction of the endemic form of leprosy, perhaps by the late Roman period,
the disease increased in prevalence, peaking during the thirteenth century or some time later.
Leprosy then declined and disappeared by the end of the fifteenth century, but remained in
Scandinavia much later than in other regions of Europe. The general pattern is similar to
treponematosis and tuberculosis in that it increases with elevation in population size, density,
and interpersonal contact. The reasons for the disappearance of leprosy remain obscure.
Increased immunities due to centuries of exposure to the disease and improvements in
hygiene and living conditions may have contributed to its decline (Carmichael, 1993; Man-
chester, 1991).
There are no distinctive sex differences in leprosy prevalence in the Næstved adults studied
by Møller-Christensen. Status differences in leprosy prevalence are difficult to determine in
this sample. Møller-Christensen did not compare possible status differences (e.g., exterior
church versus interior church burials). However, a high-status young adult male buried in the
Næstved church choir displayed classic symptoms of leprosy. Clinical evidence indicates that
leprosy can occur in anyone, but malnutrition is a predisposing factor (Keil, 1933). The
location of leprotic skeletons in peripheral areas of cemeteries associated with churches in
Sweden points to their lower status in Medieval society in this setting (Linderholm &
Kjellström, 2011). Interestingly, stable isotope analysis reveals no evidence of a difference
in either quality or type of diet in comparison with other members of the community interred
in cemeteries in this setting.

3.6 Specific infectious diseases: vectored infections


Medical historians have documented the history of vectored infections, including the great
plagues resulting in the death of millions of people in Europe during the Medieval and post-
Medieval periods (McNeill, 1998). The bioarchaeological investigation of treponematosis,
tuberculosis, and leprosy make clear the enormous burden of infectious disease on health

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
112 Exposure to infectious pathogens

and quality of life in past populations. These diseases, however, represent only a part of the
record. Historical sources potentially provide a more complete picture of past living circum-
stances and the larger picture of morbidity and mortality. For example, bubonic plague, which
arguably transformed the health and demographic landscape of Europe in the Middle Ages to
the end of the eighteenth century, has until recently been known from historical documenta-
tion alone as it leaves no pathognomonic signature in skeletal remains. Historians have
documented the progression in Europe of two pandemics of plague, assuming that these
plagues – the Justinian plague (AD 541 – 767) and the Black Death (AD 1347 – 1750) – were
bubonic plague, a rodent disease caused when humans are bitten by fleas carrying Yersinia
pestis (Carmichael, 1993; Haensch et al., 2010). However, some authorities have challenged
the notion that these pandemics were bubonic plague, arguing for example that mortality
profiles were unlike bubonic plague in the modern era and that anthrax or viral hemorrhagic
fever were more likely causes (Scott & Duncan, 2001; Shrewsbury, 1970).
Biomolecular analysis of bone from human remains recovered from plague cemeteries in a
variety of locations in Europe (Bianucci et al., 2008, 2009; Bos et al., 2011; Drancourt et al.,
2007; Garrelt & Wiechmann, 2003; Haensch et al., 2010; Harbeck et al., 2013; Kacki et al.,
2011; Raoult et al., 2000; Schuenemann et al., 2011; Tran et al., 2011; Wiechmann & Grupe,
2005), dating to the first and second pandemics, reveals the presence of DNA of Y. pestis (pla
and gplD genes). Moreover, Bianucci and her collaborators (2008, 2009; Kacki et al., 2011)
have developed a diagnostic immunological test for detecting the F1 antigen, which is
antigen-specific for Y. pestis. Where aDNA analysis was inconclusive, this test confirms the
presence of Y. pestis in a series of skeletons from individuals from various settings. In
particular, F1 antigen was found in remains having the pla and gplD genes, at Lambesc
(AD 1590 plague event) and Marseille (AD 1722 plague event). These details are essential for
understanding the causes, epidemiology, and evolution of disease processes specific to plague
(Dobson, 2007). Similarly, immunological analysis of bone samples in areas known to have
suffered from endemic malaria now reveal a protein signature for the presence of Plasmodium
falciparum, the parasite responsible for malaria across many areas of the world (Bianucci
et al., 2008; Fornaciari et al., 2010a, 2010b). This now makes possible the documentation of
another major disease having no skeletal signature.

3.7 Summary and conclusions


It cannot be overstated how important infectious diseases are in shaping the evolution and
history of our species. Unfortunately, the record of infectious disease as it is expressed in
archaeological human remains is understated owing to the fact that most deaths due to
infectious disease are not recorded in the skeleton or that the lesions due to infectious
conditions may not be diagnostic of any particular disease. However, as a result of major
advances in the development of some key diagnostic tools for pathogen identification,
especially including via histology (Mays, 2008; Schultz, 2001, 2003) and molecular analysis
in general, and DNA analysis in particular (Donoghue, 2008), there are new opportunities for
providing clearer understanding of specific diseases, their history, and the impact on health
and well-being. DNA analysis is cautiously exciting as it has the potential to confirm the
presence of such important diseases as tuberculosis, leprosy, malaria, bubonic plague, and
potentially treponematosis. Perhaps even more important than identification, however, is the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
3.7 Summary and conclusions 113

new insight into the human/pathogen co-evolution and co-adaptation. Yet, DNA analysis is
not a panacea that will solve the ambiguities surrounding more traditional difficulties of
definitively diagnosing specific diseases in bone. Some investigations underscore that ancient
pathogen DNA recovery and analysis can be quite unreliable and can be fraught with
ambiguity (Roberts & Ingham, 2008; Wilbur et al., 2009). Bioarchaeological analysis of
ancient disease may indeed be strongest when multiple lines of evidence are used in concert,
such as differential diagnoses drawn from visual inspection of bone, radiographic imaging,
and DNA analysis (Klaus, Wilbur et al., 2010). Nevertheless, the so-called “Next Generation”
DNA sequencing techniques that are currently emerging may hold the promise to overcome
many of these issues and usher in a new era in pathogen DNA analysis in bioarchaeology.
In the New World, infectious disease appears to be relatively more common in late
prehistoric settings than in earlier periods. A similar pattern is expressed in Medieval Europe
in comparison with pre-Medieval Europe. In general, these increases are linked with increased
population size and aggregation, mostly in agricultural, agropastoral, and urbanized societies.
Poor dental health (based on elevated levels of dental caries, periodontal disease, wear, and
tooth loss) is related to dietary factors, and to a lesser extent, use of the masticatory complex
for mechanically demanding, nonmasticatory functions. Skeletal infections are especially
prevalent in populations living in densely settled communities. Thus, it should come as no
surprise that treponematosis, tuberculosis, and leprosy – as well as elevated levels of non-
specific bone infections – are present in these settings. In Medieval Europe, co-occurrence of
leprosy and tuberculosis reflects a similar deterioration in living standards.
These general characteristics support the contention that infectious disease as it is
expressed in osseous remains is essentially density-dependent. These diseases are opportunis-
tic in that individuals exposed to them are already stressed by poor diets and at high risk of
early death. In at least some of these settings, poor diets likely exacerbated previously
compromised health, while increased population density and poor sanitation enhanced the
burden of infectious disease and its likelihood of widespread transmission.
The presence of skeletal indicators of infection, both nonspecific and specific, is indicative
of long-term responses to pathogens. In a sense, therefore, the lesions reflect vigorous
immune response – the individual survived the initial pathogenic attack long enough to elicit
a skeletal response (Ortner, 1991; Powell, 1988). However, presence of these lesions in high
frequencies in many of the groups discussed in this chapter also reflects an elevated disease
burden and generally negative impact on adaptation and health (compare with Goodman,
1994; Goodman & Armelagos, 1989). In this regard, inflammation is traditionally thought of
as a natural outcome of infection, helping the body to heal. However, inflammatory responses
to a wide range of infections, including dental caries and periodontal disease, promote
cardiovascular disease and elevated mortality (Li et al., 2000), and thus have a detrimental
impact on overall systemic health.
The data on infectious disease, skeletal and dental, reflect social dynamics within popula-
tions. For example, caries rate differences between males and females in many societies
indicate differential access to foods and variability in dietary behavior. Differences in patterns
of health by rank or status in some groups reflect probable differences in exposure to stressors
and quality of life generally. Regardless of how infectious disease affected specific compon-
ents of the populations, the overall impact was negative – affected individuals likely had
reduced ability to acquire key resources (e.g., food) and to participate in essential work

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
114 Exposure to infectious pathogens

activities, and may well have had shortened lifespans. Individuals affected by infectious
disease carried heavy social burdens. With regard to leprosy in Medieval Europe, the disease
was highly feared, and individuals with the disease were considered living dead, who were to
be isolated, forgotten, and removed from society (Moore, 1987).
Various causal factors discussed in this chapter underscore the fact that infectious disease
has a varied etiology. Undoubtedly, specific pathogens are linked with particular infectious
diseases, and are identified as their “causes.” However, even when hosts are infected by these
pathogens, actual disease transpires only when pathogen virulence coincides with host
susceptibility in a conducive environment.
The prevalence of skeletal lesions in an archaeological population does not show a direct
one-to-one correlation with actual prevalence in a living population. For example,
tuberculosis was highly prevalent in some pre-antibiotic groups, but it only rarely spreads
to the skeleton (Ortner, 2003), generally reported as 3% – 7% of cases in living populations.
Some archaeological series show comparatively higher prevalences (Buikstra & Williams,
1991; Eisenberg, 1986; Milner & Smith, 1990; Roberts and Buikstra, 2003). Additionally, it
is important to distinguish between disease and infection. Disease prevalence – whether drawn
from living or extinct populations – may represent only a small part of the total picture of
infection. Similarly, the risks outlined in this chapter for infection may differ considerably
from the factors that ultimately influence and determine if disease will develop from the
initial infection.
Diseases involving skeletal tissues must have contributed significantly to the burden of ill
health in many earlier societies, as they do today. Various segments of populations may have
been affected by infectious disease differently, but the experience was always mediated by
local ecological, cultural, social, and behavioral circumstances.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 12 Jan 2017 at 18:46:30, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.005
4 Injury and violence

4.1 Introduction
Human skeletal remains from archaeological contexts provide a rich and diverse record of
past behavior, reflecting the wide array of circumstances – ecological, cultural, social,
economic, and political – that put a society and its members at risk of injury and death,
resulting from either accident or violence. Many injuries are not identifiable in human
skeletons, especially those that are limited to soft tissues. Moreover, accidental death is
virtually invisible in the archaeological record except under very special circumstances, such
as in building collapse or natural disasters (Cicchitti, 1993; Deiss, 1989; Palkovich, 1980;
Sakellarakis & Sapouna-Sakellaraki, 1991). Violent death is often invisible in the study of
skeletal remains, except in instances when bone is directly damaged and shows no evidence of
healing, such as by gunshot or a sharp-edged weapon.
Despite these limitations, osteological remains are a highly useful index for assessing
outcomes relating to unintentional (accidental) and intentional, malevolent (violent) injury
in a remarkable variety of contexts around the world (Arkush & Tung, 2013; Courville, 1962;
Domett et al., 2011; Jimenez, 1994; Judd & Redfern, 2012; Knowles, 1983; Knüsel & Smith,
2014; Lambert, 2007, 2014; Lovell, 2008; Martin & Anderson, 2014; Martin & Frayer, 1997;
Martin et al., 2012; Merbs, 1989a; Ortner, 2003; Robbins Schug et al., 2012; Roberts &
Manchester, 2005; Schulting & Fibiger, 2012; Šlaus et al., 2012; Tung, 2012a, 2012b; Walker,
2001a; Webb, 1995). This applies especially because skeletal injuries provide a snapshot of a
person’s lived experience at that particular moment when s/he sustained the injury, often
resulting in death.
There is an abundance of skeletal injury data presented in the bioarchaeological literature.
In the past, the combined sparseness of a population perspective and dominance of descrip-
tion in much of this literature, however, largely precluded the realization of the enormous
potential that these kinds of data have for drawing inferences about human behavior and
conflict situations in earlier societies. However, a richer understanding of this important
record for inferences about behavior in the past has developed, especially starting with work
on the linkages between the injured and their environmental and social contexts, the two of
which are often linked (Domett et al., 2011; Judd, 2008; Judd & Redfern, 2012; Knüsel &
Smith, 2014; Martin et al., 2012; Milner et al., 1991; Nichols & Crown, 2008; Robbins Schug
et al., 2012; Schulting & Fibiger, 2012; Tung, 2012a, 2012b; Walker, 2001a). The trauma
patterns documented in skeletal remains provide an interpretive tool for understanding the
social implications of gender and power relations. Thus, bioarchaeological study of trauma
emphasizes biocultural context and the association with person, setting, and the array of
(often complex) social and behavioral circumstances leading to injury and/or death.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
116 Injury and violence

Several problems challenge the study of injury in archaeological skeletal remains, in


addition to those discussed earlier. Primary among them is the confusion that sometimes
arises between skeletal damage originating from accidental or violent causes versus damage
having nothing to do with past human behavior. For example, damage to bone produced by
shovels, trowels, and other equipment during archaeological excavation can produce marks
on bones that mimic blade-induced cutmarks (Milner et al., 2000). This confusion has led to
fanciful reconstructions of conflict and its consequences for health and well-being (Blakely &
Mathews, 1990). Other postdepositional alterations to bone that can be confused with cut-
marks and other forms of trauma are cracking and weathering, root stains, and small
carnivore and rodent damage. The application of methods developed in taphonomic and
forensic sciences adds much needed rigor to the identification of injury and violence in
archaeological human remains (Botella et al., 2000; Galloway, 1999; Hutchinson & Hum-
phrey, 2002; Kennedy, 1994; Lewis, 2008; Walker, 2001a).
Because both perimortem trauma – injury occurring at or around the time of death – and
postmortem fracture show no evidence of remodeling, the two are sometimes difficult to
distinguish. This problem is well illustrated by controversies surrounding the study of arch-
aeological skeletal remains (Constandse-Westermann, 1982; Judd, 2008; Schimmer, 1979). In
addition, the lack of healing may simply make it impossible to identify a fracture (Figure 4.1).
Various forms of skeletal trauma are identified by osteologists, ranging from self-inflicted
injuries (Tyson, 1977) to trauma involving excision of pieces of cranial vault (trephination or
trepanation) (Andrushko & Verano, 2008; Christensen & Winter, 1997; Erdal & Erdal, 2011;
Hrdlička, 1941; Liesau von Lettow-Vorbeck & Pastor Abascal, 2003; Lorkiewicz et al., 2005;
Margetts, 1967; Merbs, 1989a; Parker et al., 1985–1986; Romero, 1970; Sankhyan and Weber,
2001; Stewart, 1957; Stone & Miles, 1990; Webb, 1988, 1995; and many others), amputation or
decapitation (Bloom et al., 1995; Cybulski, 2014; Dupras et al., 2010; Friedmann, 1973; Mays,
1996; Tucker, 2014; Tung, 2012a; Verano et al., 2000), and evulsion and mutilation of teeth
(Bennike, 1985; Coppa et al., 2006; Humphrey & Bocaege, 2008; Oxenham et al., 2006; Robb,
1997a; Tayles et al., 2007; Temple et al., 2011; Turner, 2004; White et al., 1997; see review in
Milner & Larsen, 1991). Because virtually all skeletal trauma can be attributed to either accident
or violence, this chapter focuses on the behavioral interpretations of these two circumstances.

4.2 Skeletal injury and lifestyle


The general lifestyle conditions in past societies are revealed by the assessment of injuries
observed in archaeological human remains. Prevalence of skeletal injuries by element type as
well as regional temporal patterns gives insight into the influence of different lifestyles and
landscapes, such as hunting and gathering in mountainous regions or farming in river valley
flood plains. Each of these settings involves exposure to different kinds of terrain and use of
technologies that lend themselves to different risks of injury. These injuries can have serious
consequences, resulting in impaired function and reduced ability to survive.
Clinicians identify a variety of injuries relating to accidents and accidental circumstances,
such as fractures of the lower leg (tibia, fibula), clavicle, ribs, upper arm (humerus), and hip
(especially the femoral neck) (Bulger et al., 2000; Magnusun, 1942; Ortner, 2003; Stawicki
et al., 2004; Stimsom, 1943). These patterns are also present in archaeological human remains
(Ortner, 2003; Roberts & Cox, 2003; Roberts & Manchester, 2005). In comparison with other

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.2 Skeletal injury and lifestyle 117

Figure 4.1 Perimortem cranial depressed fracture at the glabella of the frontal bone;
Phum Snay, Cambodia. (Domett et al., 2011; reproduced with permission of authors and
Antiquity.) (A black and white version of this figure will appear in some format. For the
color version, please refer to the plate section.)

elements, many past human groups show relatively high prevalence of rib, vertebral, and
radial (especially distal) fractures relating to accidental injuries and hazards of living circum-
stances (Cybulski, 1992; Dickel, 1991; Gasperetti & Sheridan, 2013; Grauer & Roberts, 1996;
Judd, 2008; Kaplan et al., 1977; Molleson, 1992; Ortner, 2003; Rose, 1985; Sandzén, 1979;
Todd, 1927; Ubelaker, 1981). Injuries to these regions provide a picture of behaviors that can
be dangerous, and they show how particular lifestyles and adaptations related to either
subsistence practices or occupation can be a risk in terms of accidental injury.

4.2.1 Rib injury


In their review of the studies of fractures from more than 30 000 skeletons from 201 sites in
Britain, Roberts and Cox (2003) found that rib fractures are the most common for the time
spanning the Roman period (beginning AD 43) through the post-Medieval period (ending

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
118 Injury and violence

Figure 4.2 Antemortem fracture of rib angles from a single traumatic event; Birmingham,
United Kingdom. (Brickley, 2005; reproduced with permission of author and John Wiley &
Sons, Inc.)

AD 1850). In this context, Brickley’s (2006) analysis of an early modern (AD 1750–1850) series
from Birmingham, England, reveals that 15.6% of individuals had at least one rib fracture,
with higher prevalence in older adults, in males, and in individuals from less-wealthy graves.
The record suggests that the fractures are likely mostly due to work-related accidents
(Figure 4.2). A relatively higher prevalence in lower versus higher social classes suggests that
lower social classes were at greater risk of being injured in this way. Clinical evidence
indicates the dangers of rib fractures for health. That is, in addition to extreme levels of pain,
rib fractures are linked to increased mortality, often resulting from respiratory complications,
reduced function, and increased likelihood of infection. In the pre-antibiotic era, a rib fracture
could mean early death (Bulger et al., 2000).
The risks of urban living are also well documented in a nineteenth-century Euroamerican
series from Belleville, Ontario (Jimenez, 1994). Adult males and adult females show a very high
prevalence of rib fractures, likely arising from accidents (males¼46.6%; females¼27.9%).
As with Britain, these frequencies reflect the increased risk of injury during the period of
industrialization and the development of urban life.

4.2.2 Spondylolysis and other vertebral trauma


Spondylolysis is a degenerative pathological condition that involves a separation of the
vertebral neural arch in the area between the superior and inferior articular processes (called
the pars interarticularis) (Figure 4.3). Some researchers assume that the condition is inherited

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.2 Skeletal injury and lifestyle 119

Figure 4.3 Spondylolysis of a lumbar vertebra; Avery, Georgia. (Photograph by Mark


C. Griffin.)

(Shahriaree et al., 1979; Snow, 1974; Wiltse, 1957; Wynne-Davies & Scott, 1979). A more
compelling explanation than genetic origin is that spondylolysis is a type of fracture: it is
absent at birth, usually absent in childhood, and increases progressively through adulthood
(Bridges, 1989a; Lester & Shapiro, 1968; Merbs, 1989b, 1995; Stewart, 1953b, 1956, 1979;
Ward & Latimer, 2005; Weiss, 2009). Moreover, there is healing in later adulthood (Merbs,
1995), the separation affects only intact bone (Wiltse et al., 1975), and it develops gradually in
response to excessive mechanical loads over a period of time (Eisenstein, 1978; Roberts, 1947;
Wiltse, 1962; Wiltse et al., 1975; and see Merbs, 1983, 1989b; Weiss, 2009).
The condition is unique to humans, suggesting that bipedality plays an important role,
largely owing to lordosis as an evolutionary adaptation in order for the body’s trunk to
maintain balance over the pelvis (Bridges, 1989a; Merbs, 1989b; Ward & Latimer, 2005). The
defect is usually bilateral (Merbs, 2002; Waldron, 1992) and almost always involves the fifth
lumbar vertebra only, although it also occurs with decreasing frequency from the inferior to
the superior lumbar spine in clinical circumstances (Moreton, 1966; Roche & Rowe, 1951) and
archaeological settings (Bridges, 1989a; Gunness-Hey, 1980; Lester & Shapiro, 1968; Lundy,
1981; Mays, 2006a; Merbs, 1983, 2002; Pálfi, 1992; Pilloud & Canzonieri, 2012; Snow, 1948;
Stewart, 1979; Waldron, 1993; Wilczak et al., 2009). The defect has also been found in
cervical and thoracic vertebrae, but these are rare occurrences.
Stewart (1953b) and Lessa (2011) reported unusually high prevalences in prehistoric Aleut-
Eskimo and coastal Brazilian populations, respectively. In the latter, for example, nearly one-
third of adults were affected, with younger adults expressing a greater frequency than older
adults. In the series studied by Stewart from north of the Yukon River, Alaska, more than 40%
of individuals exhibit separate neural arches. Inherited structural characteristics of vertebrae

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
120 Injury and violence

may predispose the individual to the defect under mechanically stressful conditions. For
example, Stewart (1956) observed that spondylolysis is associated more frequently with
(1) a long “pre-arcuate” spine; (2) an acutely inclined superior sacral surface; (3) pronounced
lumbar lordosis; and (4) minimal curvature and depth of superior sacral articular facets.
Others suggest predisposing factors, including size of the articular processes (Nathan, 1959),
large pars interarticularis vascular foramina (Miles, 1975), scoliosis (McPhee & O’Brien, 1980),
presence of lumbar-sacral transition vertebrae (Merbs, 1983), spina bifida (Merbs, 1989b),
facet joint distances (Ward & Latimer, 2005), and spina bifida occulta and anterior inclination
of the sacral table (Mays, 2006a). There may well be genetically based predisposing factors for
spondylolysis, but the mechanical environment prompting its appearance is required.
High frequencies of spondylolysis in laborers (Lane, 1893) and in other individuals involved
in mechanically demanding activities, such as in college football and other athletic sports
(Hoshina, 1980; Jackson et al., 1976; McCarroll et al., 1986; Soler & Calderón, 2000), support
the mechanical stress model. The low frequency of spondylolysis in industrial populations
engaged in activities involving minimal physical exertion is consistent with this interpret-
ation. For example, one sample of twentieth-century Americans has a prevalence of only 7%
(Moreton, 1966), and the Terry and Hamann-Todd collections involve similarly low preva-
lences (Roche & Rowe, 1951; Ward & Latimer, 2005; and see Fredrickson et al., 1984).
Eighteenth- and nineteenth-century Londoners from Spitalfields also have low frequencies
that are slightly greater in males than in females (2.2% vs. 0.6%; Waldron, 1993). These values
are lower than those in earlier populations, suggesting a decrease in physical demands as it
affects the lower back in British populations (Waldron, 1991a, 1991b; and compare with
Stirland, 1996).
A range of activities involving hyperextension of the back, perhaps accompanied by jarring
and twisting, are linked with spondylolysis (Merbs, 1989b; Ward & Latimer, 2005). Given the
broad range of behaviors that are associated with these defects, no specific mechanically
related activities or stresses appear to lead to spondylolysis (Bridges, 1989a; Merbs, 1989b).
Merbs (1989b) speculates that spondylolysis may have an adaptive value in that it may
engender flexibility of the lower back (see also Snow, 1974). Bird and coworkers (1980), for
example, reported that adults with the defect considered themselves “more supple in youth”
than adults who lack the defect. The defect seems to have a minimal negative influence on
physical performance. For example, college football players with the defect appear to lose
neither practice nor playing time, and they continue to play professionally in later years (see
discussion in Merbs, 1989b).
The close association between mechanical loading of the lower back and spondylolysis
indicates that the defect should be documented in archaeological skeletal series. In
northwestern Alabama, early prehistoric hunter-gatherers have a higher prevalence of
spondylolysis than late prehistoric agriculturalists (Bridges, 1989a). Statistical analysis
of the co-occurrence of spondylolysis and osteoarthritis reveals only a weak relationship
between the two conditions. Osteoarthritis prevalence is broadly similar between individuals
with and without spondylolysis. Structural analysis of long bones reveals a general increase in
mechanical demand in the agricultural groups, which seems to contradict the findings based
on spondylolysis prevalence (see Chapter 6). Bridges (1989a) speculates that spondylolysis
may be associated with unusual postures or specific activities affecting the lower spine rather
than overall activity levels (compare with Hoshina, 1980).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.2 Skeletal injury and lifestyle 121

Adult males in populations with appreciable frequencies of spondylolysis have a higher


prevalence than adult females (Arriaza, 1995, 1997; Gunness-Hey, 1980, 1981; Mays, 2006a;
Merbs, 1983, 1995; Pilloud & Canzonieri, 2012; Stewart, 1931, 1979; Trembly, 1995; Wal-
dron, 1993; Weiss, 2009). Presumably, mechanical demands causing the defect were greater
for men than for women in these societies. There are some notable exceptions to this pattern,
which emphasizes the role of mechanical demand and the influence of culture cross-cutting
gender lines. Contact-era Omaha and Ponca women from northeastern Nebraska, for example,
have a higher prevalence of spondylolysis and spinal osteoarthritis than men (Reinhard et al.,
1994). Highly demanding activities affecting the female spine are well documented ethnohis-
torically, such as hide scraping undertaken in a stooped posture (Reinhard et al., 1994). In
addition, women were responsible for a range of other physically demanding activities not
shared by men, including house construction and firewood gathering.
Archaeological and other data provide good contextual information on why prevalence of
spondylolysis is elevated in some settings. In the Mariana archipelago, adults have a very high
prevalence of spondylolysis (approximately one-third), with males showing considerably
higher levels than females (Arriaza, 1997; Trembly, 1995). These elevated levels of spondylo-
lysis may be related to heavy lifting and carrying associated with construction of large
megalithic structures called latte. Some of the latte stones weigh hundreds of pounds, and
required a great deal of physical effort to move them. The construction of these buildings by
men explains such a striking frequency of spondylolysis in adult male skeletons (Arriaza,
1997; Trembly, 1995). Similarly, in colonial-period Spanish Florida where the native popula-
tion was exploited heavily for labor, there is a marked increase in the prevalence of spondylo-
lysis compared to the pre-colonial population, with males having a higher prevalence than
females (Larsen & Ruff, 1994).
Comparison of a rural medieval series from Wharram Percy (c. AD 1000–1400), England,
with modern western Europeans reveals a considerably higher prevalence of spondylolysis in
the former (Mays, 2006a). These findings are consistent with the kind of lifestyle and degree of
physical labor experienced in Medieval village life. That is, repetitive lifting, digging, plowing,
and other activity conducted while standing and walking would have entailed high strenuous
activity and especially heavy mechanical loading of the lower back. These kinds of physical
activity were prevalent throughout adulthood in English Medieval farmers generally (Hana-
walt, 1986).
Anterior slippage of a vertebral body relative to another immediately below it sometimes
occurs following a spondylolytic fracture (Wiltse et al., 1975). The dislocation, called spon-
dylolisthesis, is normally prevented by the restraining effects of muscles, ligaments, inter-
vertebral disks, and especially by the buttressing provided by intact vertebral articular
processes. In the absence of this support, the vertebra is displaced forward due to gravity
(Merbs, 1989b). The degree of slippage ranges from barely perceptible to complete anterior
displacement of the superior body relative to the inferior body (Merbs & Euler, 1985). In
archaeological settings, spondylolisthesis is difficult to distinguish from post-interment dis-
placement arising from other causes (e.g., rodent activity), but matching of degenerative
changes (e.g., osteophytes) on two adjacent vertebrae can be used to identify the condition
(Manchester, 1982; Merbs & Euler, 1985).
The vulnerability of the vertebral column to other types of fracture has also been observed
in the lower cervical and upper thoracic vertebrae. In a range of populations, the tip of the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
122 Injury and violence

spinous process is fractured and displays a pseudarthrosis (Knüsel et al., 1996). Clinical
evidence suggests that the fracture is due to highly forceful muscle contraction involving
muscles of the neck or activities involving scapular retraction toward the vertebral column
during rib elevation (Knüsel et al., 1996). Virtually all individuals affected, archaeological and
contemporary, are males. Although rarely reported, avulsion (disruption) injuries involving
vertebral end plates provide a potentially important perspective on spinal injuries. Unlike
spondylolysis, which involves gradual, repeated trauma leading to fracture, vertebral evul-
sions are acute injuries, such as caused today by “whiplash” trauma in automobile rear-end
collisions (Maat & Mastwijk, 2000). In these kinds of injuries, there is abrupt hyperflexion–
hyperextension. As documented in modern Holland, skeletal changes include anteriorly
displaced annular epiphyses, compression fracture, and anterior vertebral rim fracture (Maat &
Mastwijk, 2000). As with spondylolysis, males in this series have a higher prevalence than
females, strongly suggesting greater demands on the spine in males than females, likely
relating to labor involving rapid movement.

4.2.3 Skeletal injury of accidental or violent origin


Injury in modern and archaeological contexts is sometimes difficult to attribute to accidental
or violent origin. For example, so-called parry fractures – fractures involving the distal half of
the ulnar diaphysis – are common in a wide variety of human populations (Alvrus, 1999;
Angel, 1974; Bassett, 1982; Bennike, 1985; Brothwell, 1961; Burrell et al., 1986; Judd, 2008;
Kilgore et al., 1997; Lovejoy & Heiple, 1981; Møller-Christensen, 1958; Neves et al., 1999;
Smith, 1996, 1997; Smith & Jones, 1910; Stewart, 1974; Ubelaker, 1981; Webb, 1989, 1995).
As the eponym indicates, parry fractures are interpreted as resulting from the individual
attempting to ward off a blow directed at his/her head or upper body by using the forearm
(Angel, 1974; Armelagos, 1969; Judd, 2008; Jurmain, 1991; Lahren & Berryman, 1984;
Manchester, 1983; Pietrusewsky & Douglas, 1994; Salib, 1967; Smith & Jones, 1910; Webb,
1989, 1995; Wells, 1982; Wood, 1979; but see Smith, 1996, 1997) (Figure 4.4). However, their
association with Colles’ fractures in a number of settings (Smith, 1996), which are likely as a
result of injury sustained from a fall, suggests that some parry fractures may also be due to
accident. Without consideration of context, it is virtually impossible to attribute origin of the
fracture (Lovell, 1997; Walker, 2001a).
Colles’ fractures – fractures of the distal radius – have highly specific behavioral implica-
tions. The specificity of these fractures in modern settings indicates that the fracture is due to
a person attempting to break a fall by placing the hand and arm outward, landing on the
outstretched hand, with dorsal displacement of the distal radius as a result (Judd, 2008; Mays,
2006b; Silman, 2003) (Figure 4.5). Thus, Colles’ fracture largely reflects risks leading to
accident. It is possible, however, that a person can be pushed from behind, fall, and break
the fall with an extended arm, an action resulting in fracture of the distal radius.
If the forearm is used to protect the head from injury, then forearm diaphyseal fractures
(especially parry fractures) and cranial injuries should coincide. In order to demonstrate an
association between forearm and cranial injuries, Smith (1997) determined frequencies of
forearm and cranial injuries in a sample of prehistoric skeletons (n¼1695) from Tennessee. In
this series, ulnar and radial fractures occur with some frequency, but head and face injuries
are extremely rare. These findings suggest that, at least with this population, forearm fractures

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.2 Skeletal injury and lifestyle 123

Figure 4.4 “Parry” fracture of right ulna. In this case, the fracture site resulted in a disunion,
or pseudoarthrosis; Lambayeque, Peru. (Photograph by Haagen Klaus.) (A black and white
version of this figure will appear in some formats. For the color version, please refer to the
plate section.)

may have been caused by accidents and not by aggressors (Smith, 1997). Alternatively, the
populations were simply successful at fending off blows, thus protecting their heads from
injury.
Conversely, crania from the Santa Barbara Channel Islands have a very high frequency of
violence-related traumatic injuries (18.3%; 138/753; and see later) and few parry fractures
(<3%) (Lambert, 1994). Forearm fractures are equally distributed throughout the temporal
sequence, whereas cranial trauma has a distinctive peak frequency in the Early Middle period
(c. 1490 BC–AD 1150). In the southern hemisphere, at a Wari-era (AD 600–1000) site in Peru,
nearly one-third of the adults have cranial trauma and 4% of the left ulnae exhibit parry
fractures (Tung, 2007). Like the series from Tennessee, these findings suggest that the
attribution of forearm fractures to parrying is better understood in relation to broad patterns
of skeletal trauma rather than a single cause such as protection of the head.
Individuals in some settings likely did use the arm to avert blows to the head, thus resulting
in forearm injury. For example, in the Honokahua precontact sample from Hawaii (Pietru-
sewsky and Douglas, 1994), and various series from Australia (Dinning, 1949; Pretty &
Kricun, 1989; Webb, 1989, 1995), Nubia (Alvrus, 1999; Judd, 2008; Smith & Jones, 1910),

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
124 Injury and violence

Figure 4.5 Medial view of distal right radii: healed Colles’ fracture (left); normal (right);
anatomical specimens. (Photograph by Paul Braly.)

North America (Smith, 1996, 1997), and South America (Tung, 2007), left-side fractures have
a higher prevalence than right-side fractures, sometimes significantly so (e.g., Honokahua:
68.8% left vs. 31.3% right). Perhaps this reflects the victim holding up the left arm for
protection while holding an implement in the right hand in order to inflict injury on the
opponent.
Although behavioral circumstances associated with parry fractures cannot be known in all
circumstances, ethnographic and historic documentation provides some perspective suggesting
that the injured person was a victim of violence. For example, hand-to-hand fighting is well
documented in native Australians (Webb, 1989, 1995). Parrying shields were widely used for
protection against clubs (Massola, 1963). Kricun (1994) notes that parrying fractures could have
arisen either from direct blows to the forearm or from blows hitting the shield with the forces
transmitted to the ulna. These observations, coupled with the high levels of cranial trauma,
indicate the strong likelihood that parry fractures arose in conflict situations. On the other hand,
historic sources for ancient Nubia indicate the presence of both violence (warfare with
their Egyptian neighbors: Buzon & Richman, 2007) and sport (stick fighting: Carroll, 1988).
Interestingly, the relatively high frequency of parry fractures in Sudanese Nubia has consider-
able time depth, lasting from at least the Bronze Age (2500 BC) to sometime during the Meroitic
period (post-400 BC) (compare with Alvrus, 1999; Judd, 2008; Smith & Jones, 1910).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.2 Skeletal injury and lifestyle 125

4.2.4 Fracture risk: bone mass, fragility, and lifestyle


In addition to the risks described previously – terrain, interpersonal violence, occupation, and
a host of other factors leading to fracture – a paramount risk, especially for mature adults, is
age-related bone loss and the associated increased fragility (Espallargues et al., 2001; Marshall
et al., 1996). Indeed, the greatest concern surrounding bone loss in older adults today is the
increased risk of fracture following a fall or acute traumatic event. In addition to bone mass,
other risk factors include changes in chemical properties, material properties of bone, such as
accumulation of microdamage caused by repeated loads from normal activity, and structural
factors, such as reduced trabecular connectivity (Agnew & Bolte, 2012; Burr et al., 1997;
Glencross, 2011).
The study of archaeological remains provides significant insight into factors contributing to
increased risk that are not considered in the biomedical literature. Independent of bone mass
and structural and other properties, diseases affecting ambulatory function likely contribute
in important ways to increased fracture risk, especially in individuals having diseases that
impair vision and sensory perception. In the Medieval leper hospital cemetery from Chiches-
ter, England, fracture prevalence reached 15%, some three to five times greater than in non-
leper samples from contemporary sites in the region (Judd & Roberts, 1998). Judd and Roberts
(1998) make the strong case that leprous individuals have an increased risk of fracture due to
ambulatory impairments. These impairments are associated with reduced vision resulting
from the disease and deterioration of fingers and toes, which elicits general difficulty in
walking.
The link between bone loss and increased risk of fracture in archaeological contexts has
been systematically investigated only rarely, despite a large body of research dealing with
other aspects of bone mass, such as aging, functional adaptation, and growth (Ives &
Brickley, 2005; Lazenby, 2002; Mays, 2006b; Rewekant, 2001; and see Chapter 2). Building
on the profound importance of bone mass (and loss) as a risk factor in fracture patterns and
prevalence in past populations, Glencross and Agarwal (2011) have investigated the large
Neolithic series from Çatalhöyük, Turkey. The series is especially well documented bioarch-
aeologically (Hillson, Larsen et al., 2013; Larsen et al., 2013) and archaeologically (various in
Hodder, 2013), providing an excellent context from which to interpret patterns of bone loss
and other potential risk factors relating to living conditions in an early farming context.
Populations living in this setting practiced mixed foraging and farming, with significant
reliance on wheat agriculture. Various health indicators reveal a remarkably low prevalence
of morbidity, and biomechanical analysis shows very high robusticity, indicating a lifestyle
that involved considerable loading and elevated activity. In general, the changes in bone
mass (as measured by cortical mass in metacarpals), in comparing young (20–29 years),
middle (30–49 years), and older (50þ years) adults, represent respective periods of peak bone
mass, stasis and beginning bone loss, and accelerated bone loss (Glencross & Agarwal,
2011). Some interesting variation absent in modern clinical settings is well documented in
this sample. That is, early adulthood shows peak bone mass, followed by beginning of bone
loss in middle adulthood, and considerable bone loss in older adulthood. However, sex
differences at Çatalhöyük deviate from differences documented in modern populations. In
this setting, bone mass in older females and males is similar, in contrast to modern
populations whereby older females show considerably less bone mass than older males. This

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
126 Injury and violence

Figure 4.6 Antemortem diaphyseal fractures of the right radius (right) and left ulna (left);
Çatalhöyük, Turkey. (Photograph by Scott Haddow.) (A black and white version of this
figure will appear in some formats. For the color version, please refer to the plate section.)

greater loss of bone in females than in males is due to hormonal changes (estrogen
reduction) at menopause. The similarity in bone mass in comparing older Çatalhöyük
females and males in their bone mass may reflect the influence of activity and gender roles
on bone metabolism. Interestingly, with the exception of two young adults with metabolic
disruption and abnormal bone, the population shows no evidence of fracture relating to
increased bone fragility. Thus, at least in this setting, age-based increases in fracture risk are
minimal, likely owing to a similarity in bone mass. Fractures certainly exist in this setting
(Figure 4.6), but the prevalence and risk are minimal.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.2 Skeletal injury and lifestyle 127

Overall, the study of fracture prevalence in relation to age and risk in archaeological
contexts underscores the very important point that lifestyle and the lived experience are
much more complex than the clinical record (largely based on a relatively narrow spectrum of
variation in recent humans) would indicate. Like the Çatalhöyük series from Turkey, the
Libben series from eastern North America expresses a very different pattern than in recent
urban populations (Lovejoy & Heiple, 1981). Analysis of the number of years at risk reveals
that fracture rates peaked in two age groups, adolescence/young adulthood (15–25 years) and
old adulthood (>45 years). This pattern may reflect an elevated risk of trauma due to warfare
and conflict, especially for the younger group. Lovejoy and Heiple (1981) note that women
and men were equally affected by injury, indicating that the elevated rates were due to
accidents associated with activities such as hunting forays and travel (in adults) or play (in
juveniles).

4.2.5 Fracture and accident


The presence of occasional fractures and injury is documented in the fossil record, but the
record of trauma does not stand out until the increased abundance of skeletal fractures and
other injuries in late archaic Homo sapiens (Neanderthals) in the later Pleistocene. Traumatic
lesions resulting mostly from accidental injury have been found in European and western
Asian Neanderthals since the first upper limb fracture among these groups was reported in the
mid-nineteenth century (Schaaffhausen, 1858; and see Berger & Trinkaus, 1995; Estabrook &
Frayer, 2014). Virtually every relatively complete adult Neanderthal skeleton older than 25–30
years displays some type of injury (Berger & Trinkaus, 1995; Gardner & Smith, 2006),
including those from violence.
The Shanidar sample presents a highly distinctive picture of injury in late Pleistocene
hominins. Several of the six adult skeletons display skeletal trauma, mostly of accidental
origin (Stewart, 1977; Trinkaus & Zimmerman, 1982). Shanidar 1, a mature adult male, has
multiple injuries. The bones of the right upper arm and shoulder are less than half the size of
the bones of the left upper arm and shoulder (Figure 4.7), which may have been caused by
either childhood nerve damage or adult disuse atrophy following a severe injury (Trinkaus &
Zimmerman, 1982). The diminutive right humerus has two fractures, both at the distal end.
One of the fractures may be one side of a false joint (pseudarthrosis) or evidence of an
amputation (Stewart, 1977; Trinkaus & Zimmerman, 1982). This individual also displays
evidence of severe cranial trauma involving extensive scarring of the frontal bone and a
crushing fracture of the left eye orbit. All postcranial and cranial injuries were completely
healed at the time of death.
Other accidental trauma includes a rib fracture in Shanidar 4, a mature adult male. The
presence of a large callus at the fracture site, but with exposed trabecular bone, indicates that
healing was incomplete at the time of death. Finally, like Shanidar 1, another adult male
(Shanidar 5) has frontal bone trauma. The injury was completely healed at death.
The Shanidar cranial injuries are part of an overall pattern of head and neck injuries in
European and western Asian late archaic Homo sapiens (Berger & Trinkaus, 1995). Nearly
one-third of these hominins have head and neck trauma, which is more than twice the
prevalence of a clinical sample from New York (Berger & Trinkaus, 1995) (Table 4.1).
A survey of traumas associated with a variety of occupations in recent humans indicates that

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
128 Injury and violence

Figure 4.7 Atrophy and healed fracture of the right humerus; Shanidar, Iraq. The left
humerus is normal. (Photograph and copyright by Erik Trinkaus; reproduced with
permission.) (A black and white version of this figure will appear in some formats. For the
color version, please refer to the plate section.)

American rodeo athletes also have a high prevalence of head and neck injuries relative to
other regions of the body (Berger & Trinkaus, 1995). The pattern in rodeo athletes reflects the
dangers of riding highly irritated animals (e.g., Bos taurus, Equus caballus); namely, head and
neck injuries in rodeo athletes result from impacts after being thrown from these animals. By
logical extension, the high prevalence of head and neck injuries in Neanderthals relates to

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.2 Skeletal injury and lifestyle 129

Table 4.1 Distributional frequencies (%) of traumatic lesions by anatomical region:


Neanderthals, recent archaeological samples (Bt-5, Libben, Nubia), clinical samples (London,
New York, New Mexico), and rodeo athletes (Adapted from Berger & Trinkaus, 1995: Table 2)
Head/neck Trunk Shoulder/arm Hand Pelvis Leg Foot

Group
Neanderthal (n¼17) 29.6 14.8 25.9 3.7 3.7 11.1 11.1
Bt-5 (n¼223) 1.8 51.1 22.4 6.3 3.1 9.0 6.3
Libben (n¼94) 6.4 21.3 29.7 0.0 0.0 39.4 3.2
Nubia (n¼160) 10.6 6.9 53.1 1.9 3.8 22.6 1.3
London (n¼1730) 6.2 7.0 31.6 24.4 0.2 23.6 7.0
New York (n¼11 959) 13.7 12.3 25.3 21.9 0.5 20.6 5.6
New Mexico (n¼792) 1.6 12.5 23.1 23.6 2.1 11.1 25.9
Rodeo (n¼181) 39.2 9.9 25.9 6.1 3.3 6.1 9.4

hunting activities, especially involving close encounters with medium-sized ungulates (Berger
and Trinkaus, 1995). Use of spears would have necessarily placed hunters in close proximity
to, and hence, bodily injury from, their enraged prey.
Angel (1974) attempted to document temporal trends by comparing post-Mesolithic arch-
aeological skeletons (n¼2125) and modern Euroamerican samples. This comparison reveals
several general characteristics of accidental injury: adult males have more injuries than adult
females; older adult females have more fractures than older adult males (Buhr & Cooke, 1959);
and recent populations have far more fractures than earlier populations. The higher frequency
of skeletal injury in adult males than females in Angel’s samples likely reflects a greater
exposure of men to trauma. The reversal in older adults likely indicates the effects of bone
loss, especially in postmenopausal women, and especially in modern industrial populations
where lifestyles are distinctively different from traditional societies sampled from the arch-
aeological past (Glencross & Agarwal, 2011). Angel (1974) contends that the higher fracture
prevalence in the modern samples results from the hazards associated with twentieth-century
technology and urban living, such as the reliance on automobiles, walking on staircases, and
urban crowding.
In addition to the anatomical region analysis discussed earlier, analysis of fracture preva-
lence for the skeleton can provide insights into local adaptation, such as those associated with
particular subsistence patterns. Contrary to Steinbock’s (1976) assertion that trauma
decreased with the foraging-to-farming transition, there is considerable variation in patterns
and prevalence of fracture that are largely driven by local circumstances (various in Cohen &
Armelagos, 1984; Cohen & Crane-Kramer, 2007; Steckel & Rose, 2002). This record suggests
that while injury prevalence owing to accidents may have decreased with the rise of farming
in some regions, there are no clear temporal generalizations that can be made about injury
relating to accidents. Nevertheless, the recent analyses of injury viewed in temporal perspec-
tive in specific communities and regions provide critical perspective on adaptation, lifestyle,
and risk.
Some North American prehistoric maize agriculturalists show generally low prevalence of
accidental injuries. At Moundville, for example, the total frequency of bones affected is only
0.7% (Powell, 1988). Many of these fractures are associated with lower-status individuals; no

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
130 Injury and violence

high-status adult females have accident-related traumatic injuries (Powell, 1988). Thus, at
least in this setting, sedentary populations had a relatively accident-free lifestyle, and high-
status, elite women may have been spared altogether from activities resulting in accidental
injury (lack of elite males in the skeletal series precluded their assessment). In contrast, at the
late prehistoric Mississippian site of Chucalissa, Tennessee, high-status males have far higher
frequencies of fractures than low-status males or high-status females (Lahren & Berryman,
1984). Unlike the Moundville population, most of the injuries can be attributed to violence. In
contrast, late prehistoric intensive agriculturalists from the Dickson Mounds site show an
increase in fracture prevalence compared with earlier less-intensive agriculturalists (Good-
man, Lallo et al., 1984). This suggests that injury risk increased with the shift to more
intensive farming in this setting of the American Midwest.
In prehistoric Thailand, Domett and Tayles (2006) tested the hypothesis that more estab-
lished agriculturalists in this setting would be exposed to increased risk of injury from
extensive land clearance and other farming activities. Comparisons of Neolithic farmers with
later Bronze Age intensive farmers reveal a 10-fold temporal increase in fractures, from 0.3%
to 3.0%. Some of the fractures are parry and craniofacial, suggesting some evidence for
increased violence and competition for resources. The finding of increased skeletal injury in
the more intensive Bronze Age farmers in Thailand is consistent with results comparing urban
and rural populations in Medieval Britain. That is, the farming population from Raunds shows
four times the frequency of fracture than the urban samples (19.4% vs. <10%) (Judd &
Roberts, 1999; Roberts & Manchester, 2005). The breadth and risk of activities of farming in
Medieval England – such as plowing, tree felling, herding, transport of products – likely had
considerably more risk than village life where individuals were more tethered to the home and
craft specialization.

4.3 Intentional injury and interpersonal violence


Violence has always played a role in social relations, ranging from physical confrontation
between two individuals to large-scale warfare between nations. This record of violence and
conflict is represented by archaeological evidence such as fortifications, defensible site
locations, settlement pattern, weaponry, and iconographic and symbolic representations
involving weapons, places, and people in conflict (Arkush, 2011; Domett et al., 2011; Haas &
Creamer, 1993; Keeley, 1996; Larson, 1972; Lichtheim, 1973; Maschner, 1992; Redmond,
1994; Schulman, 1982; Steponaitis, 1991; Tung, 2012a; Wilson, 1987; and many others).
These site characteristics usually only identify the threat of conflict and not its actual
outcome. For example, in precontact Alabama, fortifications appear with the adoption of
maize agriculture, yet the bioarchaeological record of violence shows a decline in conflict in
later prehistory (Danforth et al., 2007).
Ethnographic observation can provide an important source of information on violence and
aggression in human societies (Burbank, 1994; Ember & Ember, 1997; Ferguson, 1995; Haas,
1990; Harrod et al., 2012; Montagu, 1978; Redmond, 1994). Harmony and cooperativeness are
sometimes emphasized by anthropologists for many nonliterate societies, which is in sharp
contrast to reality (Berndt, 1978; Ember, 1978; Erchak, 1996; Fienup-Riordan, 1994; Keeley,
1996; Lambert, 2002c; Walker, 2001a). Skeletal injuries represent clear testimony to conflicts
between once-living individuals. Thus, archaeological skeletons are regarded as the only

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 131

direct evidence of violent interaction. Viewed in light of living conditions and social,
economic, and political context in archaeological settings, much can be learned about the
origins and use of violence in past populations.
The skeletal and paleopathological literature on violence and injury is dominated largely by
descriptions of limited samples, such as individual instances of arrow wounds (Armendariz
et al., 1994; Bovee & Owsley, 1994; Lewis & Lewis, 1961; Pryor, 1976), decapitation (Bennike,
1985; Harman et al., 1981; Jones, 1908; McKinley, 1993; Newman & Snow, 1942; Smith,
1993; Ubelaker, 1988; Wakely & Bruce, 1989; Webb, 1974;), dismemberment (Brothwell,
1971; Smith, 1993; Snow & Fitzpatrick, 1989; Webb, 1974), sacrifice and ritual violence
(Bennike et al., 1986; Fowler, 1984; Kanz & Grossschmidt, 2006; Klaus, Centurion et al., 2010;
Pijoan & Mansilla, 1997; Stead et al., 1986), mutilation, especially scalping (Allen et al., 1985;
Hamperl & Laughlin, 1959; Hoyme & Bass, 1962; Lesley, 1995; Miller, 1994; Neumann, 1940;
Ortner, 2003; Ortner et al., 2008; O’Shea & Bridges, 1989; Smith, 1995), and cranial depressed
fractures or other forms of head injury (Haas & Creamer, 1993; Lux, 1994; Manchester &
Elmhirst, 1980; Wenham, 1989; and many others). Because many of these studies usually
involve the investigation of one or several skeletons only, they frequently provide limited
information for inferring conflict behavior in the populations from which they were drawn.
However, placement of these few skeletons within their larger, contextualized settings where
various lines of evidence about society at a particular place and time offers important insights
into circumstances and causes surrounding violent encounters (Arkush & Tung, 2013; Cook,
2012; Judd & Redfern, 2012; Klaus, 2012; Milner, 1999; Milner et al., 1991; Montgomery and
Perry, 2012; Tucker, 2014; Tung, 2012a, 2012b; Walker et al., 2011).
Contextually oriented approaches in bioarchaeology are revealing the fundamental import-
ance of skeletal data for documenting patterns of violent behavior in the past, ranging from
interpersonal conflicts between two individuals to full-scale warfare involving scores of
participants. The strength of these approaches is their reliance on context, derived from the
clinical and epidemiological literature on modern people, ethnography, climate history,
ethnohistorical accounts, and archaeology for interpreting violence, its origins, its conse-
quences, and its distinctive skeletal signatures. Indeed, skeletal trauma is among the most
common pathological conditions in the bioarchaeological record. The interpretation of the
evidence of injury requires a detailed series of interpretive pathways by the investigator
(Figure 4.8). Reflecting this level of attention paid to interpreting trauma in archaeological
contexts, there is a growing record of insight into the history of violence going back many
thousands of years.
In the following discussion, a series of bioarchaeological studies are assessed that collect-
ively illustrate the enormous variation in skeletal evidence of conflict in past populations. This
discussion is intended to be neither comprehensive nor synthetic; rather, a representative
sample of various kinds of skeletal injury useful for identification of patterns of conflict is
discussed. These studies are drawn from a diverse set of geographical and cultural settings,
including the American Midwest (Norris Farms, Riviere aux Vase), Great Plains (McCutchan-
McLaughlin, Crow Creek, Larson Village), and Southeast (Koger’s Island, Chickamauga),
southern California Pacific coast (Santa Barbara Channel islands), Arctic (Kodiak Island,
Admiralty Island, Saunaktuk), American Southwest (Anasazi), Latin America (multiple local-
ities), Easter Island, Sudan (Nile Valley), across the continent of Australia, and Southeast Asia
(Cambodia). The study of human remains from these settings provides important perspectives

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
132 Injury and violence

Injury occurred yes no A perimortem injury or


Does the injury show
before death postmortem damage
signs of healing?

Injury occurred yes


Is break consistent with
around time of death fresh-bone fracture?

Inference:
Post-depositional damage
during site formation?

Did the injury occur no A perimortem injury or


before death? occurred after death

yes
Is there active yes A perimortem injury or
new bone formation? occurred before death

no

Inference: Inference: Inference:


Injury was not immediately fatal, Possibly associated with Injury possibly associated with
possible implications for efficacy the cause of death, mortuary rituals, cannibalism,
of medical practices and social suggests accidental injury or other postmortem
support system or homicidal violence processing by humans or
scavengers

Figure 4.8 Model for interpreting traumatic lesions on skeletons. (Walker, 2001a;
reproduced with permission of Annual Reviews Inc.)

on violence, chiefly in nonliterate tribal and state societies, mostly before contact with
Western populations.
Lastly, patterns of northern European violence, ranging from interpersonal conflict to
large-scale preindustrial warfare and execution, and the subsequent spread of European forms
of violence to the New World and military campaigns in North America, are addressed. These
investigations illustrate variability of traumatic injury in skeletal remains as well as the
impact of violence on different components of the populations involved (e.g., gender, age,
status) within a specific period or consequent to major adaptive shifts.

4.3.1 American Midwest


Norris Farms
The study of skeletons recovered from a cemetery near a late prehistoric Oneota culture (c. AD
1300) village on bluffs overlooking the Illinois River floodplain provides a comprehensive
picture of widespread violence and conflict in a prehistoric tribal society (Milner, 1995, 1999;
Milner & Smith, 1989, 1990; Milner et al., 1991; Santure, 1990). Sixteen percent of the
264 skeletons have evidence indicating violent death, mostly in the form of unhealed trauma.
The range of unhealed trauma affecting primarily the cranium, body trunk, and upper limbs is
striking, especially regarding projectile wounds, holes in crania produced by stone celts,
depressed fractures, and various forms of mutilation. Some individuals have multiple skeletal
injuries, far exceeding what would have caused their deaths. Location of traumatic injuries in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 133

Figure 4.9 Cut marks on adult frontal bone (scalping); Norris Farms, Illinois. (From Milner &
Smith, 1990; reproduced with permission of authors and Illinois State Museum.) (A black
and white version of this figure will appear in some formats. For the color version, please
refer to the plate section.)

individual skeletons indicate that some were attacked from behind and some from frontal
assault. The former shows that some individuals attempted to flee. Evidence for mutilation is
especially abundant in the Norris Farms series. Multiple perimortem cutmarks on cranial
vaults (especially on frontal bones) produced by stone tools reveal that at least 14 individuals
had been scalped (Figure 4.9). Individuals missing their skulls and having cut cervical vertebra
show that they had been decapitated. Presence of cutmarks on multiple articular regions of
postcranial bones evinces widespread dismemberment of limbs. Many instances of punctures
and gouges in skeletal elements, mostly produced by carnivores, suggest that a significant
proportion of victims (n¼30) were exposed above ground for a period of time before
interment.
Active bone infections, articular joint dislocations, and partially healed bone fractures and
other trauma denote the presence of severe and long-standing disabilities for many individ-
uals at the time of their deaths. These conditions may well have impaired their ability to
escape confrontation, leading to early death (Milner, 1995; Milner et al., 1991).
The presence of victims – individuals who clearly exhibit evidence of violent trauma – in
individual burial pits or pits containing a few individuals indicates that violence occurred over
the entire period of the use of the cemetery, which spanned at least several decades. Over this
timespan, not all of the victims died outright as a result of violent encounters. For example, a
number of adults, both non-survivors and survivors of attack, display embedded chert
projectile points (Figure 4.10a,b).
The pattern of deadly conflict at Norris Farms is similar to that of ethnographic small-scale
societies where violence is endemic (Chagnon, 1992; Harrod et al., 2012; Heider, 1979;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
134 Injury and violence

(a)

Figure 4.10a Arrow point in adult sternum; Norris Farms, Illinois. (From Milner & Smith,
1990; reproduced with permission of authors and Illinois State Museum.) (A black and
white version of this figure will appear in some formats. For the color version, please refer
to the plate section.)

Schmidt & Schröder, 2001). Raids in these societies often involve ambush and surprise attacks,
but may also occur during chance encounters. Victims of attacks in these groups can include
individuals of all ages and both sexes. At Norris Farms, fully one-third of the adults were
victims of violent attack at one point in their lives, and they include equal numbers of adult
males and females; only two are juveniles (<15 years of age). The equal number of adult
females and males with traumatic injuries is different from ethnographically observed small-
scale societies where males are the predominant victims (Chagnon, 1992; Divale & Harris,
1976; Harrod et al., 2012; Heider, 1979; Keeley, 1996; and others). Female captives provide

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 135

(b)

Figure 4.10b Projectile point in left tibia; Norris Farms, Illinois. (From Ryan & Milner, 2006;
reproduced with permission of authors and Elsevier Ltd.) (A black and white version of this
figure will appear in some formats. For the color version, please refer to the plate section.)

economic return – women’s labor in many societies is essential for food collection and
preparation. Relatively, more women may have been killed at Norris Farms because of their
burden on the attacker’s resources, or it may simply have been too much trouble for the
attackers to bring captives back to their home village (Milner, 1995; Milner et al., 1991).
Warfare at Norris Farms is tied to the highly dynamic sociopolitical circumstances that
characterize this region of the American Midwest during later prehistory. The Oneota represent
an intrusion into the central Illinois River valley, replacing a somewhat more organizationally
complex system (Mississippian). Clear evidence of social tensions is indicated by fortifications
and defensible settlement locations here and throughout the late prehistoric Midwest during a
period of population expansion and growing tensions over and competition for resources
(Milner et al., 2013). Populations occupying the region may have been in competition over
productive lands and resources concentrated in river valleys (Milner et al., 1991). Violence in
this case may have been a strategy employed to gain control of these highly valued resources.

Riviere aux Vase


At the late prehistoric (AD 1000–1300) Riviere aux Vase site in southern Michigan, nonlethal
cranial injuries, consisting of round or elliptical vault depressions produced by wood or stone
clubs, have been identified (Wilkinson & Van Wagenen, 1993). A higher frequency of adult

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
136 Injury and violence

females (n¼14) than males (n¼5) with cranial injuries suggests that women were the preferred
target of violence. For some of these women, violent encounters may have occurred on more
than one occasion – five female crania have multiple healed depressed fractures. One of these
individuals has a severe depressed fracture and an accompanying large incision on the
occipital bone as well as multiple fractures on the left and right parietal bones.
Although other women may have been responsible for the cranial injuries in this population
(compare with Burbank, 1994), the demographic characteristics of the injured group suggest
that males were the primary aggressors. Males and females show a very different age pattern
of injury: females’ peak age of trauma is the early adult years, while male trauma is evenly
distributed across age groups. Male trauma is oriented toward the front of the vault, and
female trauma is distributed throughout the cranium. Wilkinson and Van Wagenen (1993)
suggest that violence directed at women by men or women against women or co-wives in
polygamous societies is consistent with ethnohistoric accounts of violence in American
Midwestern native populations (Wilkinson & Van Wagenen, 1993).

4.3.2 American Great Plains


McCutchan-McLaughlin
A large proportion of individuals (19%) recovered from this archaeological site in south-
eastern Oklahoma are from a single multiple burial containing the remains of nine individuals
(three infants, four adult females, two adult males) (Powell & Rogers, 1980). None of the
victims displays evidence of mutilation – such as scalping or dismemberment – that suggests
violent confrontation. However, some members of the population clearly died in a violent
fashion: large projectiles had penetrated thoracic and pelvic cavities, vertebral columns, and
limbs. These associations underscore the importance of the archaeological context for docu-
menting violence (Buzon, 2012). Unfortunately, this contextual information is easily lost if
not properly recorded during the course of excavation. Had this information not been
available, the only evidence for violence in this series would have been multiple-individual
burials, which by itself is only circumstantial evidence of violent death.

Crow Creek
The study of human remains from various late prehistoric sites in the Missouri River valley
located in the present states of South Dakota and North Dakota reveals evidence for violent
confrontations between tribal groups competing for overlapping resources and territory
(Bovee & Owsley, 1994; Hollimon & Owsley, 1994). The proto-Arikara (c. AD 1325) skeletal
series from the Crow Creek site supplies important details on prehistoric conflict (Gregg &
Gregg, 1987; Willey, 1990; Willey & Emerson, 1993; Zimmerman et al., 1981). Analysis of
some 500 individuals buried in a single pit indicates that nearly all members of a village were
massacred during the course of a single raid. The presence of carnivore tooth marks and
weathering reveals that following the attack, the deceased were exposed for a period of time
prior to their burial by returning survivors or allies (compare with Milner & Smith, 1989).
Analysis of the human remains suggests that although all of the deceased were victims of a
single attack, violence had a well-established history in the Crow Creek villagers; a significant
number of massacre victims had healed violence-related injuries, including scalping (Willey,
1990; Willey & Emerson, 1993).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 137

Virtually all individuals in the series have unhealed cutmarks on frontal bones and other
cranial elements indicating removal of the scalp with a stone knife at or following the time of
death. Forty percent of the victims have cranial depression fractures, mostly located on the top
and sides of the vaults. In addition to scalping trauma, various other mutilations are common,
including tooth evulsions, alveolar and tooth fracture, nose removal, decapitation, and
dismemberment of both upper and lower limbs. The presence of sharp-force trauma on
mandibles – especially on ascending and/or inferior borders of rami – suggests that mutilation
also involved tongue excisions.
Demographic assessment of the Crow Creek population indicates that young adult males
(15–24 years of age) outnumber young adult females by a factor of two. Additionally, there
are twice the number of older adult females (45–59 years of age) than older adult males. The
absence of young women may reflect captive taking, escape, or actual demographic compos-
ition of the population from which victims are drawn (Willey & Emerson, 1993). Similarly, the
paucity of elderly men reflects their escape or actual demographic composition. It is unlikely
that older males were captured because they would have represented a relatively low eco-
nomic return for the raiders. Perhaps older males may be missing owing to earlier raids and
endemic warfare (Willey & Emerson, 1993).

Larson Village
During the historic period, the Arikara occupied a series of temporally successive villages as
they migrated northward up the Missouri River valley. One such village was decimated by
violence. At the seventeenth-century Larson Village site in northern South Dakota, 71 skel-
etons from house floors display evidence of violent death and mutilation (Owsley et al., 1977).
About one-third (34%) of the victims had been scalped. The mutilation patterns are similar to
those displayed in the Crow Creek population, including dismemberment, decapitation, and
facial, dental, and cranial vault trauma, as well as tongue excision. Skeletal modifications on
one young adult female run the gamut of violent injury and mutilation: “the left side of her
skull had been broken away, though cuts on the frontal, right parietal and occipital indicated
she was scalped. The distal diaphysis of her right radius and ulna are articulated in anatomical
position. . .A knife cut was made through the soft tissues in order to free the hand as a trophy
or possibly to secure a bracelet. Epiphyseal areas of both bones were broken. After severing
the muscles and tendons, the assailant simply broke the hand free. Other bones have been cut,
including five right ribs, the ventral and posterior surface of the right clavicle, the right
scapula, the right tibia and both femora. A deep cut near the distal epiphysis of the right
humerus must have resulted while separating the upper and lower arm. Cuts on the femurs are
on the neck; the objective may have been to remove the legs from the body. Bones of both legs
are associated with the skeleton though neither was articulated when excavated” (Owsley
et al., 1977:125). This analysis indicates that the Larson Village was attacked, and the villagers
attempted to defend themselves in their individual houses (Bamforth, 1994). The presence of
numerous unburied remains indicates that the Larson Village ceased to exist following the
conclusion of the attack.
Demographic composition of the historic Larson Village site and prehistoric Crow Creek
massacre victims are similar; both series contain fewer young adult females and juveniles
than young adult males. This pattern suggests that children and women may have been

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
138 Injury and violence

captured rather than killed in the attack, which is documented historically in the region
(Lowie, 1954).
Analysis of scalping patterns reveals changes in warfare prior to and during the contact
period in the northern Plains (Owsley, 1994). Comparison of a large sample of late prehistoric,
protohistoric, and early historic crania (n¼751) from 15 archaeological sites indicates that
throughout the period, both males and females were at equal risk of being scalped. This risk
greatly increased for men but not for women in the early historic period; instances of scalping
tripled for males and halved for females. Most of the male victims are young adults (20–34
years of age), which likely reflects the deaths of warriors who were killed either during raids or
in defense of home villages (Owsley, 1994). Owsley’s (1994) analysis establishes that death
from violence was present throughout the late prehistoric and historic occupation of the
Plains. This violence was occasionally punctuated by eruptions of large-scale warfare
resulting in numerous deaths at one time (e.g., Crow Creek and Larson Village massacres).
Thus, contrary to earlier assertions (Newcomb, 1950), analysis of skeletal evidence of conflict
discloses that warfare and inter-group conflict did not originate during the early period of
contact with Europeans or Euroamericans, but was a well-established part of the cultural and
social behavioral repertoire of prehistoric societies living in the region. The increase in Arikara
male deaths during the historic period likely reflects elevated frequency of confrontations,
especially with the encroaching Sioux from the east (Owsley, 1994). The overall similarity
between precontact (e.g., Crow Creek) and postcontact (Larson Village) Arikara skeletal
injuries indicates an enduring pattern of conflict in this region. In the precontact Plains
setting, conflict appears to have been triggered by food shortages and stress generally, as is
suggested by the presence of stress indicators in human remains as well as paleoclimatological
evidence for periodic droughts after AD 1250 (Bamforth, 1994). Archaeological evidence also
suggests that new populations were migrating into the region during this time, resulting in
increased competition for productive lands. In view of these new developments, Bamforth
(1994) contends that the stage was set for increased violence in later prehistory, a pattern
which was exacerbated by the spread of Sioux into the Missouri River valley.

4.3.3 American Southeast


Koger’s Island
Koger’s Island is one among many archaeological localities in the American Southeast where
human remains have been recovered, mostly excavated in the mid-twentieth century, and
have contributed to an enormous bioarchaeological record (Larsen, 2012). Owing to this large
and rich data set, analysis of skeletal remains has provided an abundance of data on variation
within communities and across the region. Analysis of skeletal remains and their archaeo-
logical context provides important evidence of conflict in a late prehistoric (Mississippian
period) population from the Tennessee River valley and associated tributaries and northern
Alabama and eastern Tennessee. At Koger’s Island in the western Tennessee Valley (Alabama),
Bridges and coworkers (Bridges, 1996; Bridges, Jacobi et al., 2000; Jacobi, 2007) documented
cutmarks on the frontal and occipital bones of five adult male crania from four mass burials.
The cutmarks represent scalping and decapitation. In addition, crushing fractures on a
manubrium and a scapular spine of two individuals and deep cuts on the ribs of a third
indicate violent death. Demographic assessment of victims suggests that the impact of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 139

20

Males

15 Females
Number of Individuals

10

0
10 20 30 40 50
Age (years)
Figure 4.11 Sex-specific mortality profiles (age versus number of individuals); Koger’s
Island, Alabama. (Adapted from Bridges, 1996; reproduced with permission of
John Wiley & Sons, Ltd.)

violence in this society may have been profound – some 21% of the total number of
individuals were recovered from multiple-interment (mass) graves. The population contains
a relatively large number of infants (about 25% of the total), a pattern that is characteristic of
many preindustrial, farming-dependent populations (Bocquet-Appel et al., 2008; Lambert,
2009). Adult (>15 years of age) males and females have very different age-at-death profiles
(Figure 4.11). Female deaths peak slightly during the late teens and early twenties, progres-
sively fall to a low point in the thirties, and again rise in the forties. Male deaths are few in
number during the late teens, but high in the thirties. Presumably, many of these deaths
resulted from violent intergroup encounters. Age composition of the small burials (containing
fewer than five individuals) is different from age composition of the multiple burials (five or
more individuals). The former is dominated by infants and younger juveniles. The latter has an
unusual peak in the thirties, and most of these deaths are adult males. The pattern is
distinctive in that there are relatively few women or children in the multiple burials, suggest-
ing that women and children had either escaped or were captured; males may have died while
protecting the village from aggressors.
The loss of adults, and especially men who were responsible for protection of the group and
acquisition of resources not acquired by women (e.g., animal protein), would have had far-
reaching consequences for the population’s ability to mitigate stress in a hostile setting. In
later prehistory, political systems and population size declined in this and some other regions
of the American Southeast and Midwest (Anderson, 1994; Milner, 1990; Steponaitis, 1991).
This hostile environment may have contributed to the decreased presence of these groups in
later prehistory (Bridges, 1996).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
140 Injury and violence

Comparisons of the Koger’s Island sample with other late prehistoric populations in the
region suggest that conflict was highly localized. Less than 1% of the late prehistoric
(Mississippian period) Moundville skeletons (n¼564) from west-central Alabama have skeletal
injuries, and only a handful of these are from violence (one piercing wound and three with
cuts) (Powell, 1988). Similarly, upriver in the eastern Tennessee River valley, Mississippian-
period populations display relatively few violence-related injuries (Smith, 2003). Populations
from the preceding Late Woodland period in the Tombigbee River valley display an abun-
dance of injuries deriving from violence (Cole et al., 1982; Hill, 1981; Welch, 1990). Some of
the deceased who had died violently were buried in multiple-individual graves. Thus, in late
prehistoric west-central Alabama, there was an apparent decline in violence from the Late
Woodland to Mississippian periods, which Steponaitis (1991) argues was brought about by
increasing control of the regional population by more centralized polities, such as that at
Moundville. This pattern of declining interpersonal violence is certainly not a general out-
come, however. For example, analysis of the Cumberland River valley region in Tennessee
reveals a significant presence of injuries, including scalping, blunt-force trauma, and other
injuries in the Mississippian period (Worne et al., 2012). In general, however, there is an
overall spike in violence in later prehistory in Eastern North America (Milner et al., 2013), but
prevalence is highly variable.

4.3.4 Southern California Pacific coast


Aboriginal populations of the western Pacific coast of North America are often characterized
as passive and non-warlike. Early Spanish accounts highlight the peaceful nature of these
groups (Bolton, 1927; Kroeber, 1925; Priestley, 1937). Historically, eruptions of warfare and
violence during the contact period have been attributed to disruption of the natural harmony
of the region by invading Europeans (Lambert, 1994; Walker, 1989).
The study of an extensive skeletal series of prehistoric human remains from the Santa
Barbara Channel islands and mainland illustrates the inaccuracies of the perceptions of early
explorers. Walker and Lambert (Lambert, 1994, 1997, 2007; Lambert & Walker, 1991; Walker,
1989; Walker & Lambert, 1989; Walker et al., 1989) analyzed skeletal injuries – mostly
depressed cranial vault fractures and projectile wounds – in more than 1 700 individuals over
a 7 500-year temporal span of prehistory. Cranial trauma is quite common: 18.3% of
753 crania have depressed fractures (Lambert, 1994) (Figure 4.12). Most fractures are healed,
indicating that the majority of individuals survived the violent encounters. Apparently, the
intention of the individual wielding the weapon was to injure – rather than kill – the targeted
victim. It is also possible, however, that the victims successfully defended themselves from
what was intended to be a lethal blow.
Demographic analysis of the victims reveals a number of tendencies. Very few of the
victims are under the age of 10; adolescents have three times the number of vault injuries
than younger individuals. Depressed fractures are common in adults, but they are especially
common in individuals under 40 years of age. This age-specific pattern suggests that adults
not involved in behaviors and activities resulting in cranial trauma had somewhat greater
longevity in comparison with victims (Lambert, 1994).
Most of the injured adults are males. In these individuals, roughly two-thirds of the injuries
occur on the left side of the frontal, indicating that conflicts were face-to-face encounters

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 141

(a)

(b) (c)

(d) (e)

Figure 4.12 Depressed fractures in adult crania: ellipsoidal parietal injury (a), deep
circular injury (b), multiple circular injuries (c), ellipsoidal occipital injury (d), circular
parietal injury (e; arrowhead points to residual fracture line); Santa Barbara Channel
Islands, California. (From Walker, 1989; reproduced with permission of John Wiley &
Sons, Inc.)

with a right-handed perpetrator striking the left side of the head of the victim (Lambert, 1994).
This patterning of nonlethal trauma is remarkably similar to trauma observed in the Yano-
mami foragers of Venezuela (Chagnon, 1992). Yanomami men attack other males with heavy
wooden clubs. Although numerous casualties result from these encounters, they are rarely
lethal.
Adult female cranial injuries are haphazardly distributed on the face and vault – only about
one-third of the injuries are on the frontal bone (Lambert, 1994). This random distribution of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
142 Injury and violence

depressed fractures indicates that although females were occasionally involved in face-to-face
attacks, most were from other directions, including from behind (e.g., while fleeing from an
attacker). Additionally, some of these nonfrontal injuries could be from accidental causes
(Walker, 1989).
The presence of projectile injuries and associated projectile points in some individuals
reveals appreciable numbers of deaths caused by violence: of 1744 individuals, 3.3% had
been killed or wounded by single or multiple projectile injuries. Unlike the cranial trauma,
the majority of projectile victims (at least 71%) died from their wounds, indicating the
lethal intentions of the attacker. The demographic composition of individuals with either
lethal (projectile wounds) or nonlethal (cranial depressed fractures) injuries is similar in at
least two respects: first, young and mature adult males are the most affected; and second,
children, older adults, and women are the least affected. Like cranial trauma, aggression
leading to injury and death involved primarily adult males under 40 years of age
(Lambert, 1997), which is consistent with patterns observed for many nonindustrial
communities globally (Chagnon, 1992; Meggitt, 1977; and see discussion in Lambert,
1994).
The temporal patterns of nonlethal and lethal skeletal injury are distinctive in prehistoric
southern California coastal populations. Nonlethal cranial depressed fractures are common
in the Early Middle period (1490 BC–AD 580), whereas lethal projectile injuries are far more
common in the Late Middle period (AD 580–1380) (Lambert, 1994, 1997). The increase in
frequency of projectile injuries may be tied to the adoption of the bow-and-arrow in
California during the sixth century AD (Lambert & Walker, 1991). Perhaps the rapid
adoption of the bow-and-arrow at this time may have been fostered by competition and
conflict between populations in North America generally (Blitz, 1988). Regardless of the
motivation for increased lethal violence, the later peak in projectile injuries signifies a shift
to more serious – and deadlier – forms of conflict in later prehistoric southern California
populations.
The availability of abundant bioarchaeological, archaeological, and climatological data
provides a comprehensive understanding of factors that may have motivated violent behav-
ior in the Santa Barbara Channel region. Analysis of tree ring and other climatological data
suggests that the Middle period saw an increase in environmental instability, periodic
droughts, and decreased terrestrial resource productivity. These changes, coupled with
warming of the Pacific Ocean during this time, reduced marine productivity (Lambert,
1994, 1997, 2007) along with increased population size and territoriality (Kennett, 2005).
In light of these changes, Lambert (1994) proposes that elevated competition from increased
resource stress may have engendered an increase in violence in these populations. This
hypothesis is consistent with other skeletal data showing a decline in quality of life and
increase in health stress (Lambert, 1993, 1994, 2002c; Walker, 1986; Walker & Lambert,
1989; and see Chapters 2 and 3). Along with the increase in trauma and disease during this
time, there is evidence for increasing social complexity. For example, burial of high-status
grave goods (e.g., beads) with infants is more common during the peak in drought in later
prehistory, suggesting ascribed rather than achieved social rank. Perhaps increased inter-
group competition – and, hence, increased violence – for scarce resources during episodes of
environmental degradation fostered more complex social organization in later prehistory
(Fischman, 1996).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 143

Far from being a peaceful setting of the New World, the analysis of violence in the Santa
Barbara Channel island region reveals the presence of endemic violence, peaking during key
periods of resource stress and sociopolitical change. Similarly, analyses of a large bioarch-
aeological database of 16 820 burials from 329 archaeological series from central California
(San Francisco Bay area, Central Valley, and Sierra Nevada foothills) documents clear
patterning of three forms of violence-related trauma, including trophy-taking/perimortem-
dismemberment and sharp-force and blunt-force trauma (Pilloud et al., 2014; Schwitalla
et al., 2014). At least half of the sites represented in the central California database express
some form of violence. Evidence for all three forms of conflict-related injury is well repre-
sented in the region and in highly patterned ways. With regard to perimortem dismember-
ment, there is a low, but significant presence, in contrast to southern California where it is
completely absent (Andrushko et al., 2005, 2010; Bertelink et al., 2014). This form of violence
peaked during the Early Middle period (500 BC–AD 420) and then largely disappeared. Blunt-
force trauma is also well represented, with males and older adults having the highest
prevalence in general (Figure 4.13). This form of trauma peaked very late in the sequence,
being mostly limited to the eighteenth and nineteenth centuries when there was a four-fold
increase in males and two-fold increase in females. The highest prevalence during this period
was in regions that were contested by multiple tribes following European encroachment. The
pattern of cranial trauma shows that the peak is considerably later in central California
(historic period) than southern California (Early Middle period, 1490 BC–AD 580). Moreover,
unlike the nearly 20% prevalence in southern California (Walker, 1989), the prevalence in
central California is well below 10%.
Sharp-force trauma (from projectiles) is also well represented in the central California
region, but with a different pattern than blunt-force trauma. In this regard, sharp-force
trauma is present in 7% of individuals, and is much more common in males (11%) than in
females (5%). The relative prevalence for females is higher in adolescent and younger
individuals. Like blunt-force trauma, the peak frequency of sharp-force trauma is in the
historic period, with the peak being especially higher for males (16%). Overall, while males
are clearly more often involved in violent encounters, females were also engaged in confron-
tations, not just as recipients, but as combatants. The historical records and ethnographic
observations document a greater involvement of males in conflict, but also document the
important role of women, including female aggression directed at other females (Goldschmidt
et al., 1939).

4.3.5 The Arctic and Subarctic


As in California, native groups living in the Arctic are often perceived as living in a state of
quiet repose and nonviolence (Fienup-Riordan, 1994). Early documentation of Eskimos often
presents them as “passive to the point of lethargy” (Fienup-Riordan, 1994:322). Several
bioarchaeological investigations are providing new data that indicate the need for revision
of these perceptions. Findings from these studies suggest that although violence may have
been rare, it erupted on occasion and had dire consequences for some groups. Archaeological
analysis in the region, especially of settlement characteristics, village location, and defense
strategy, clearly indicates the concern for defense in the region in later prehistory (Maschner &
Reedy-Maschner, 1998).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
144 Injury and violence

0.04
0.03
Trophy Taking Males
0.02
0.01

Females
0.00
0.16

Males
Proportion of Burials Exhibiting

0.12
Blunt Force Trauma
0.08
0.04

Females
0.15
Sharp Force Trauma
0.10

Males
0.05

Females
0.00

Early Early Middle Late Middle Middle/Late Late Protohistoric/historic


(3050 - 500 BC) (500 BC – AD 420) (AD 420 - 1010) (AD 1010 - 1390) (AD 1390 - 1720) (AD 1720 - 1899)

Figure 4.13 Graphs depicting the relative frequency of trophy taking, blunt-force trauma,
and sharp-force trauma in adult females and males from Early to Protohistoric/historic
central California. (Schwitalla et al., 2014; reproduced with permission of authors and
Elsevier Ltd.)

Saunaktuk
The study of human remains from the Saunaktuk site located east of the Mackenzie Delta in
the Canadian Northwest Territories, Canada, provides compelling evidence for violent con-
frontation between native groups (Melbye & Fairgrieve, 1994; Walker, 1990). Human remains
from this locality represent a minimum of 35 Inuit Eskimo villagers, some of whom died

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 145

violently in the late fourteenth century AD. Evidence for violent death and body treatment in
this Arctic setting is indicated by extensive perimortem skeletal modifications, including knife
cuts, slashing, piercing, gouging, and splitting of long bones (Melbye & Fairgrieve, 1994).
None of the remains were purposefully buried. Most of the victims are juveniles (68.6%),
suggesting that adults – particularly males – may have been away hunting, leaving a
relatively defenseless group vulnerable to attack.
Hundreds of knife cuts, especially around articular joints and the neck (e.g., occipital
condyles and upper cervical vertebrae), indicate the practice of dismemberment and decapi-
tation. Numerous cuts on facial bones on most victims identify widespread facial mutilation
or disfigurement. Many other cuts, such as on clavicles and scapulae, reflect an overall pattern
associated with purposeful dismemberment, removal of muscle and other soft tissues, and
intentional mutilation. Unique to the Saunaktuk skeletal series is the presence of gouges at the
ends of long bones. Adult distal femora from two individuals display large perimortem
mediolateral gouges passing completely through the cortical and cancellous tissue. These
modifications are consistent with oral tradition describing a type of torture whereby the
victim’s knees were pierced and the individual was dragged around the village by a cord
passed through these perforations.
With few exceptions, long bones had been split into scores of longitudinal sections. The
surfaces of split bones are smooth and display tiny step fractures identical to bone modifica-
tions in butchered animal bones. Presumably, this breakage pattern was done in order to
extract the nutritionally rich marrow for consumption (Melbye & Fairgrieve, 1994; and
compare with White, 1992). The striking similarity between butchered animal remains found
in archaeological sites and the Saunaktuk human skeletal assemblage suggests that the
deceased had been cannibalized. In summary, at least some members of this late prehistoric
Saunaktuk population had been tortured, and all were murdered, mutilated, and cannibalized.
There is a rich historical record that provides a context for understanding violence between
native groups living in this region of the Arctic. In areas where Inuit and Native Americans
came into frequent contact – such as the Mackenzie Delta region – violent interactions
between the groups were commonplace. Oral traditions and historical accounts describe the
horrific nature of inter-group violence. For example, Samuel Hearne, who explored the region
in the late eighteenth century for the Hudson Bay Company, observed the murder and
mutilation of Inuit villagers by a group of Native Americans. He noted that “the brutish
manner in which these savages used the bodies they had so cruelly bereaved of life was so
shocking, that it would be indecent to describe it. . .” (Hearne, 1971:155; quoted in Melbye &
Fairgrieve, 1994:73).
The massacre at Saunaktuk may not have been an isolated occurrence in the Arctic region.
Preliminary evidence from Admiralty Island, Alaska, indicates the presence of a small number
of broken long bones and ribs as well as perimortem cutmarks produced by stone tools (Irish
et al., 1993). Although the evidence is limited, the patterns of modification are similar to those
identified from Saunaktuk (Melbye & Fairgrieve, 1994) as well as other sites in North America
where cannibalism was probably present (American Southwest: Redmond, 1994; Turner,
1993a; White, 1992; and see the following discussion).
On Kodiak Island, Alaska, skeletal modifications suggestive of dismemberment and canni-
balism have been described. At Uyak Bay, Hrdlička (1944) briefly documented a series of
scattered human remains, some of which had been broken in a manner which he interpreted

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
146 Injury and violence

as representing a practice of cannibalism. Analysis of remains from the Uyak and Crag Point
sites shows the presence of culturally modified remains of women and children, but not men
(Simon, 1992; Simon & Steffian, 1994; Steffian & Simon, 1994; and see Urcid, 1994). This
study reveals a small subsample of remains displaying cutmarks from disarticulation and
defleshing, drill holes, perimortem breakage, and longitudinal fracturing. Thus, cannibalism
was likely present in precontact Koniak Island native groups, albeit probably not at the levels
envisioned by Hrdlička.

4.3.6 American Southwest and Mexico


Anasazi
The Anasazi were one of several complexes of late prehistoric societies who were ancestral to
some of the modern native populations currently living in the “Four Corners” region (present-
day states of Utah, Colorado, Arizona, and New Mexico) of the American Southwest. The region
has been the focus of intensive archaeological investigation for more than a century, producing
an abundance of human remains. Most human remains are intentional interments, ranging
from isolated burials found in house floors, trash and storage areas, and large cemeteries.
Burials are generally singular and are accompanied by grave goods (Turner, 1993a).
A small – but highly visible – proportion of graves are multiple-individual interments
containing many disarticulated and broken skeletal elements (reviewed in Hurlbut, 2000;
White, 1992; and see Kuckelman et al., 2002; Redmond, 1994; Turner, 1993a; Turner &
Turner, 1995). Patterns of these skeletal assemblages contrast sharply with most other burials
in two major respects: (1) they are composed of unburied bone masses found on the floors of
structures or in the fill of kivas or rooms, or (2) they are clusters of human remains in pits
(Turner, 1993a). These bone concentrations almost always lack grave goods, and they are
frequently found in small and isolated sites lacking defensive constructions (Turner, 1993a;
Turner & Turner, 1995). The sites with these unusual remains are late prehistoric (c. AD
900–1650), they contain fewer than 30 individuals, and remains often are represented by
equal proportions of juveniles and adults and adult males and females (but see Martin et al.,
1995). As a general rule, no non-human faunal remains are present in these contexts. The
skeletal assemblages contain overwhelming evidence of perimortem trauma, disarticulation,
defleshing, and burning.
The tremendous amount of bone breakage in these assemblages prevents determination of
cause of death, and specifically, whether violence was a factor leading to the deaths of
individuals. Violence may very well have been involved (Turner, 1993a). For example,
alveolar bone and tooth sockets are highly traumatized, and at least one juvenile displays a
massive cranial injury.
The most thoroughly studied skeletal assemblage from the American Southwest is the
Mancos Canyon series (White, 1992). White’s meticulous investigation reveals a pattern of
skeletal modification similar to other sites from the Southwest (Kuckelman et al., 2002; Martin
et al., 2001; Stodder et al., 2010; Turner, 1983, 1993a; Turner & Morris, 1970; Turner &
Turner, 1992, 1995, 1999). The sample includes a minimum of 17 young adults and 12
children, all of whom display human-induced tool marks on their remains from defleshing,
percussion, chopping, and disarticulation. Thermal modification is widespread in the assem-
blage. Similar to the Saunaktuk series, Mancos Canyon long bones show extensive reduction

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 147

Figure 4.14 Fractured and longitudinally split humeri; Mancos Canyon, Colorado. (Tim
D. White. Prehistoric Cannibalism at Mancos 5MTUMR-2346. © 1992 Princeton University
Press. Reprinted by permission of Princeton University Press.)

and longitudinal fracturing (Figure 4.14). For example, humeri shafts are highly fragmented,
which was accomplished by hammerstone percussion and anvil fracturing (White, 1992:238).
These patterns of modification are similar to fracture patterns of animal bones resulting from
removal of flesh, disarticulation, and marrow extraction (White, 1992). The strong similarity
between bone modifications associated with processing of humans and mammals in the
region indicates that cannibalism was likely practiced at Mancos Canyon and other sites in
the region.
White (1992) notes the well-documented evidence of warfare and violence in the ethno-
historical and archaeological record of the American Southwest, including defensive sites,
intentionally burned habitations, as well as remains of deceased whose deaths had been
violent (Haas & Creamer, 1993). Based on the evidence – including lack of projectile injuries,
the pattern of body modification, and the extreme reduction of skeletal materials found at
the site – he contends that it is not possible to determine the reasons for cannibalism.
Several possible scenarios emerge, such as ritualized cannibalism involving killing and
eating of enemies, or perhaps the Mancos Canyon population engaged in culinary
cannibalism – the population was so starved that they consumed friends, associates, and
relatives (White, 1992). Well-documented historical cases of starvation cannibalism provide
important support for the latter model (Bonassie, 1989). The American Southwest can be
characterized as a marginal environment where seasonal cycles and resource variability led
to frequent episodes of crop failure and famine. Paleopathological analysis provides sub-
stantial evidence for nutritional stress in the prehistoric American Southwest (Lambert,
2014; Martin et al., 1991; Stodder, 1994; Stodder et al., 2010). These data alone, however,
do not provide sufficient evidence for why cannibalism occurred. In summary, although

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
148 Injury and violence

starvation is a plausible motivation for cannibalistic behavior, it cannot be identified as a


primary factor based on the evidence at hand.
Skeletal assemblages in the La Plata Valley, New Mexico, show the presence of a number of
characteristics displayed by the Mancos Canyon series, including breakage, burning, cut-
marks, and other alterations of crania, ribs, and lower limb bones (Martin et al., 2001). At site
LA 37592, remains of mostly juveniles and some adults found in trash deposits in the fill of a
kiva dating to the Pueblo III period (AD 1125–1300) possess perimortem modifications (cut-
marks, burning, breakage). Stable isotopic analysis of diet and documentation of pathological
conditions (e.g., porotic hyperostosis) do not distinguish the culturally modified sample from
the other skeletal assemblages lacking evidence of human-induced modifications. At least in
this setting, therefore, there are no apparent underlying differences explaining why some
individuals were treated differently than others at or shortly after the time of death.
Martin and coworkers (2001) suggest that although perimortem skeletal modifications and
similarity of cutmark patterns with butchering of animals may be associated with cannibalism
(compare with Boulestin et al., 2009; DeGusta, 2000; Hurlbut, 2000; Knüsel & Outram, 2006;
Turner, 1993a; Turner & Turner, 1999; White, 1992) in the La Plata Valley, other explanations
should be carefully evaluated. For example, they note that killing of individuals identified as
“witches” in the American Southwest is well documented historically, and appears to have
climaxed during episodes of food shortages and epidemics (Armelagos, 2008; Dongoske et al.,
2000; Ogilvie & Hilton, 2000; Walker, 2008). Alternatively, warfare and conflict during late
prehistory are well documented. Other investigations of skeletal remains where warfare has
been observed certainly indicate extensive modifications of skeletons, reflecting violent
encounters.
Evidence from Cowboy Wash in southwestern Colorado where seven individuals had been
processed in preparation for cooking and later consumption provides important support for
cannibalism, at least for this locality (Billman, 2008; Billman et al., 2000; Lambert, 2013;
Lambert et al., 2000; Marlar et al., 2000) (Figure 4.15). That is, presence of human myoglobin
in a human coprolite deposited at the same time that the human remains had been cooked
indicates the very strong likelihood that the person who deposited the coprolite had consumed
human flesh. The presence of human myoglobin and hemoglobin on groundstone implements
used for food preparation at Sacred Ridge, Colorado, similarly speaks to the strong likelihood
of cannibalism (Stodder et al., 2010).

4.3.7 Precontact and Contact Latin America


Precontact violence, ritual, and body processing
Ritualized violence and offering of remains – such as heads and other body parts – as a mediator
between the living, their ancestors, and the supernatural world has been documented in a wide
range of settings in precontact Latin America (Arkush & Tung, 2013; Benson & B oone, 1984;
Carneiro, 1990; Fowler, 1984; Klaus, 2014a; Redmond, 1994; Tung, 2008; various in Tiesler &
Cucina, 2007; Verano, 1995, 2001, 2007, 2014). Numerous iconographic descriptions and
ethnohistorical accounts provide information on sacrificial death and mutilation, while evi-
dence of perimortem body processing, and discoveries of missing heads and limbs in archaeo-
logical settings provide verification for this behavior (Conlee, 2007; Fowler, 1984).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 149

(a) (b)

(c)

Figure 4.15 Cut marks on the tibial and femoral shafts (a) and subtrochanteric region of left
femur (b) are indicative of tool use in processing. Perimortem fracturing of the parietal bone
(arrow) was observed on a burned cranium (c); Cowboy Wash, Colorado. (Billman et al.,
2000; reproduced with permission of authors and the Society for American Archaeology.)

Among the world’s best archaeological records of body processing is from a number of
settings in Latin America, especially in the Andean region. However, the earliest osteological
evidence is from the Lagoa Santa, Brazil, a region where a remarkable bioarchaeological
record of Paleoamerican presence has been documented (Neves et al., 2007). At the site of
Lapa do Santo (8540 yBP), an isolated skull and cervical vertebra display cutmarks consistent
with purposeful removal of the head. In particular, the right articular process of the sixth
cervical vertebra and the right posteroinferior mandibular ramus displays distinctive cutmarks
produced with a stone tool (Strauss et al., 2013). The context for decapitation is unknown,
whether as part of the mortuary ritual following death or associated with the death of the
individual.
The rise of complex societies in South America, including the iconic Andean empires, is
associated with an expansive record of body processing, best described in general as

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
150 Injury and violence

Figure 4.16 Pacatnamú mass burial, second layer of human skeletal remains; Jequetepeque
Valley, Peru. (Verano, 2008; reproduced with permission of author and Springer-Verlag.)
(A black and white version of this figure will appear in some formats. For the color version,
please refer to the plate section.)

mutilation, a form of treatment occurring at and associated with death. On the northern coast
of Peru, a series of mutilated human remains and contextual data from the site of Pacatnamú
(AD 1100–1350) in the Jequetepeque River valley are well documented (Faulkner, 1986; Rea,
1986; Verano, 1986, 1995, 2007, 2008). Located outside of the entrance of the primary
ceremonial complex, the remains of 14 adolescent and young adult males (mean age¼21
years) were recovered from the bottom of a defensive trench in three superimposed groups
(Figure 4.16). The groups were separated by erosional deposits, indicating that the deceased
had been deposited in the trench in three different burial episodes. Some skeletal remains for
each of the three groups show evidence of weathering, indicating that burial did not follow
death immediately. Delayed burial is also indicated by the presence of pupal cases of muscoid
flies representing different stages of insect growth in an open environment (Faulkner, 1986).
In the top-most group, injuries include multiple stab wounds (perforations on vertebrae and
ribs) sustained from different directions. This variable pattern of wound orientation suggests
that more than one individual may have been involved in the stabbing of the victim. In the
middle group, stab wounds are not evident. The presence of cutmarks on upper cervical
vertebrae indicates throat-slashing and decapitation. The bottom-most group also shows
evidence of decapitation. Five individuals from the middle and lower groups have bisected
manubria and fractured ribs, indicating that the chest had been opened forcibly. Collectively,
these perimortem traumas present a scenario of violent death and mutilation. Many of the
victims also have healed injuries (e.g., rib fractures, depressed cranial fractures) (Verano, 1986,
2008). Based on the age distribution, sex, and pattern of healed and unhealed injuries, Verano
(1986, 2008) speculates that the victims were war prisoners. This conclusion is well supported
by iconographic depictions in art from the region (Chimú and Moche) showing ritual mutila-
tion and sacrifice of war prisoners (Sutter & Cortez, 2005; Verano, 2005, 2014).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 151

35%

30%

25%
Percentage

20%

15%

10%

5%

0%
Archaic Formative Final EIP MH MH/LIP LIP LIP/LH LH
(N = 311) (N = 52) Formative (N = 377) (N = 575) transition (N = 799) (N = 555) (N = 598)
(N = 114) (N = 103)

Figure 4.17 Frequencies of adult cranial trauma throughout Andean series (EIP, Early
Intermediate Period; MH, Middle Horizon; LIP, Late Intermediate Period; LH, Late Horizon).
(Arkush & Tung, 2013; reproduced with permission of authors and Kluwer Academic/
Plenum Press.)

The links between warfare and ritual violence are well documented in the Peruvian Andes
when the Wari empire rose to its fullest influence in the region (AD 600–1000). In addition to
expansionist strategies that included engineering superiority and ideological domination, the
Wari state has long been thought to have used violence to bring populations under its control
(Lumbreras, 1974). Tung’s (2007, 2012a, 2012b) test of this hypothesis shows that imperial
core populations engaged in head-taking and the display of war trophies (see later), and that
core and peripheral populations have high rates of cranial trauma (25%–30%). For the region
overall, these high rates are part of a generally increasing frequency of violence where the
attacker targets the victim’s head (Figure 4.17). Although frequencies between the core and
periphery are similar, the lethality and head wound locations differ between the sites. At the
peripheral site of Beringa, which was located on a defensible mesa, adults have a higher
frequency of posterior cranial trauma, a pattern consistent with individuals being injured
while fleeing during a raid. They also show the greatest proportion of lethal trauma, suggest-
ing that the attackers went into the conflict with the intention of killing their opponent (Tung,
2007, 2012a, 2012b). It would appear that their concerns for defense were well founded. At the
nearby site of La Real, males show significantly more cranial trauma than females, and the
vast majority of fractures are on the anterior of men’s skulls. All cranial trauma is healed.
Importantly, 70% of those anterior wounds are on the left side, suggesting face-to-face
conflict and a hit from a right-handed attacker. This is similar to what Walker (1989) and
Lambert (1994) reported for the Santa Barbara Channel Islands populations (see earlier). The
parallels in cranial fracture patterning, combined with ethnographic evidence of ritualized
conflict resolution in the Andes and nearby Amazonia, led Tung (2007, 2012a, 2012b) to
suggest that many La Real males were injured in physical conflict resolutions, like those

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
152 Injury and violence

practiced by the Yanomami, the Oro-Wari of Amazonia (Conklin, 2001), and perhaps the
Chumash of the Santa Barbara Channel Islands.
Early accounts report on the practice of removal and curation of heads in native popula-
tions in the Andes and Amazonia, Ecuador, which continued well into the twentieth century
in some Amazonian groups (Verano, 1995). The use of decapitated heads is well represented in
Andean iconography, especially among the Nasca in south-central Peru (Verano, 1985) and
the Wari in the nearby highland Andes (Tung, 2008, 2012a, 2012b). Young adult males were
the favored victims, suggesting that they represent enemy combatants (Verano, 1995) or
perhaps local warriors. The presence of cutmarks on some skulls indicates that following
death, soft tissue had been removed with a sharp implement. Although variable in their
treatment, two features characterize trophy heads from the Nasca region: the frontal bone is
perforated with a single hole and the base of the skull is damaged, ranging from enlargement
of the foramen magnum to removal of most of the cranial base. The frontal bone perforation
was used to support a rope handle. The foramen magnum may have been enlarged for removal
of brain tissue, perhaps as a part of the mortuary ritual.
At the site of Conchopata in the Wari imperial heartland, nearly half of the trophy heads
had trauma, including one individual with a perimortem fracture indicating that the victim
received a blow to the head about the time of death (Tung, 2008). Many of the skulls show
cutmarks reflecting removal of flesh in transforming the head from victim to trophy (Tung,
2008). Although there has long been debate about whether trophy heads are from local
venerated ancestors or foreign enemies, strontium isotope analysis of the Wari trophy heads
shows that the vast majority are non-local, suggesting that the trophy heads represent enemy
victims (Tung & Knudson, 2011).
Ritual and sacrifice are well represented in several settings, especially in the Middle Sicán
culture (c. 900 – 1100), the descendant culture of the earlier Moche, in the Lambayque Valley,
northern Peru (Klaus, 2014a; Klaus, Centurion et al., 2010; Shimada et al., 2004). In this
setting, a rich body of evidence for sacrifice from the capital of Sicán and the more rural
temple site of Cerro Cerrillos is revealed from the study of hundreds of skeletons. The location
of cutmarks on anterior body surfaces of cervical vertebrae reveals that the victims had their
throats slit with sharp metal (bronze) tools (Figure 4.18). Some of the victims show cutmarks
in the anterior thoracic region for opening of the chest. Comparisons with earlier sites in the
Lambayeque Valley reveal that various forms of sacrifice and body treatment were new,
especially relating to the practice of opening the chest to extract the heart and of violent child
sacrifice (Klaus, Centurion et al., 2010).
In precontact Mexico, skeletal evidence has accrued on body processing, including cut-
marks, decapitation, dismemberment, burning, and intentional bone breakage in a limited
number of different temporal and cultural contexts (Pijoan & Mansilla, 1990, 1997). As has
been discussed for the American Southwest, cutmarks and other cultural modifications of
skeletal remains, by themselves, do not provide information on the motivations for body
processing (White, 1986). However, a rich historical and iconographic record indicates that
ritual was the key motivation for cannibalism in early contact-period Mexico.
A comprehensive picture of ritualized violence and death is well documented from human
remains recovered from Teotihuacan, Mexico. Although skeletal evidence for ritual sacrifice
has been known since the early twentieth century (Hrdlička, 1910a), the number of individuals
involved were too small to develop a record of broader implications for the practice. However,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 153

(a)

(b)

(c)

(d)

Figure 4.18 Composite distribution of cut marks associated with ritual killing events:
anterior aspect of cervical vertebrae (a), proximal clavicle (b), superolateral manubrium
(c), and sternal rib end (d); Cerro Cerrillos, Peru. (Klaus, Centurion et al., 2010; photographs
by Sam Scholes; reproduced with permission of the authors and Antiquity.) (A black and
white version of this figure will appear in some formats. For the color version, please refer
to the plate section.)

victims represented by single and multiple interments of mostly adults from the Temple of the
Feathered Serpent dating to the early occupation of the site and construction of the temple
(c. AD 100–250) have been analyzed (Cabrera Castro, 1993; Cabrera Castro et al., 1991; Serrano
Sánchez, 1993; Sugiyama, 1989). Sacrificial victims associated with the temple number about
120 individuals (Cabrera Castro, 1993). The skeletal remains have not been systematically
studied for cutmarks or other skeletal modifications. However, the position of the hands behind
the backs for some individuals suggests that they had been bound at the time of death.
Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
154 Injury and violence

The general lack of disarticulation as well as the inclusion of commingled individuals in single
burial pits (e.g., burial 190 included 18 adult males in a linear pit at the midpoint between the
southeast and southwest corners of the temple) suggests that these were simultaneous deaths
(Sugiyama, 1989). The social identity of the victims is not known. Nevertheless, in reference to
one of the multiple burials on the margins of the temple (burial 190), the inclusion of
numerous obsidian points, blades, and other offertory items suggests that they “were military
men, priestly soldiers, or men disguised as military personages” (Sugiyama, 1989:98).
Ritualized violence, sacrifice, and related behaviors are well documented in the iconog-
raphy of numerous precontact Latin American cultures. Mass burials, mutilation, decapi-
tation, and presence of trophy heads in archaeological sites in Peru and Mexico indicate that
scenes of these activities in various art forms in these regions were not just mythical events,
but rather, historically situated cultural practices.

Non-ritualized violence
Undoubtedly, the combining of ritual with violence was a highly developed social behavior in
Latin America, especially in state societies where public display of authority took on increas-
ing importance. However, like other areas discussed, the record of violence relating to
resource stress is also prominent in a number of settings. For example, temporal analysis of
remains from San Pedro de Atacama, Chile, reveals a clear pattern of elevated violence-related
cranial trauma (depressed and other types of fractures), doubling in frequency from 5%–11%
in the post-AD 600 occupation with increased sedentism and population density (Torres-Rouff
& Costa Junqueira, 2006). These increases coincided with elevated frequency of stress indica-
tors, such as enamel hypoplasia and porotic hyperostosis. Paradoxically, this period of
heightened resource stress, nutritional inadequacy, and violence was also a period of
increased prosperity, as is indicated by an increase in the quantity and quality of grave goods.
Interestingly, elite males from this setting have fewer violent-related cranial injuries than
non-elite males. These differences coincide with an apparent growth of social inequality,
particularly as it relates to the identity of injured (lower-status) victims and others (higher-
status) (Torres-Rouff, 2011). Like other settings in the New World, violence in this setting is
strongly associated with cultural and environmental circumstances, resulting in increased
competition for limited resources and social disruption generally. Moreover, level of disrup-
tion was likely partitioned along class lines.
Much of the evidence of violence in Latin America, however, relates to competition and
expansion of political control. Much of this record suggests that violence was one strategy
used by the dominant expanding polity. In the capital region of Cuzco of the Inca Empire, a
trend toward increasing cranial trauma is likely due to warfare, increasing from 2.5%
prevalence in the Middle Horizon (c. AD 600–1000) and in the Late Intermediate period
(c. AD 1000–1400) to 7.8% in the Inca Imperial period (AD 1400–1532) when the empire
reached its peak control (the area encompassing modern Peru and parts of Ecuador, Bolivia,
Chile, and Argentina) (Andrushko & Torres, 2011). Interestingly, all of the traumatic injuries
appear to be in peripheral areas and not the core region of the empire around Cuzco. The
record suggests that violence was used to subjugate and expand into peripheral areas.
Moreover, as is so well documented in historical accounts, violence did not decline in the
region during the Spanish conquest of the Inca Empire. Indeed, the study of juvenile and adult
remains shows an increase in bioarchaeological evidence of violence and traumatic injury

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 155

(Gaither & Murphy, 2012; Murphy et al., 2010). While some of the increase may have involved
internecine conflicts among native populations, almost certainly some originated from Span-
iards during this period of unprecedented population disruption and wide-scale suppression of
native communities.

4.3.8 Easter Island


The easternmost island of Polynesia, Easter Island (Rapa Nui), is among the most remote
inhabitable land masses on earth. The island was settled around AD 400 by a small founding
population. By the sixteenth century, population density peaked to an estimated 7000 to 9000
individuals living on just 180 km2 of land (Kirch, 1984; Van Tilburg, 1995). Ethnographic,
historical, and archaeological evidence indicates that warfare and conflict were fostered by
environmental degradation, soil depletion, depletion of fuel (wood), reduced food sources, and
overpopulation (Kirch, 1984; Owsley et al., 1994). Folklore documents a state of perpetual
warfare accompanied by numerous deaths, enslavement, murder, and cannibalism in this
highly circumscribed setting. Contact and interaction with Europeans during the early period
of exploration and conquest of this region of the Pacific (AD 1722–1868) seem to have
exacerbated the deteriorating conditions, resulting in even more violent confrontations
between groups as well as between natives and Europeans.
The study of human remains provides important data on health, lifestyle, and evidence of
violence (Gill & Owsley, 1993; Owsley et al., 1994). Based on counts of frontal bones, 11.4%
(31/271) of Easter Island adult (>15 years of age) crania have depressed fractures. Other
cranial bones also exhibit fractures, primarily the left and right parietal bones. These findings
indicate a high frequency of violent interactions, specifically involving face-to-face encoun-
ters. Adult males have roughly twice the number of cranial injuries of the adult females. More
young adults (15–34 years of age: 13.2% of frontal bone injuries) are affected than older
adults (>35 years of age: 10.3% of frontal bone trauma), suggesting the violence may have led
to early death. Most depressed fractures are either circular or oval, and were probably caused
by the impact of either hand-held rocks or clubs. In addition, fragments of obsidian blades
imbedded in some of the cranial fractures indicate the use of other types of weapons.
Extensive healing indicates that very few of these injuries were fatal, which suggests that
violence was generally not intended to have a lethal outcome (compare with Lambert, 1994,
2002c; Walker, 1989; and authors cited earlier). The arrival of Europeans resulted in the
introduction of firearms weaponry, which may have contributed to a shifting of these
intentions from nonlethal injury to homicide. For example, small lead pellets from a gunshot
wound were imbedded in the frontal and left parietal bones of a historic-period adult. The
bone tissue surrounding the entry wounds is fully healed, suggesting that the victim survived
for a period of time following the attack. Most other victims of firearms were likely not as
fortunate as this individual.
In summary, the skeletal evidence of trauma on Easter Island is only partially consistent
with folklore documenting frequent and violent interpersonal conflict, a perception for much
of the Pacific but not especially well documented bioarchaeologically for this vast area of the
world (DeGusta, 2000; Scott & Buckley, 2010). Contrary to this record, bioarchaeological
analysis indicates that violence resulted in nonlethal injury rather than widespread death
(Owsley et al., 1994). Additionally, the practice of cannibalism is not confirmed by the study

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
156 Injury and violence

of skeletal remains. Thus, folklore of Eastern Islanders overemphasizes warfare, violence, and
cannibalism.

4.3.9 Nubia and the Nile Valley


Conflicts in the Nile Valley – the setting of Egypt and Sudan – during the age of the rise of
complex societies are a matter of legend. One of the best bioarchaeological records is from
Sudanese Nubia, representing a range of localities and times (Alvrus, 1999; Buzon & Judd,
2008; Buzon & Richman, 2007; Judd, 2002, 2004, 2006; Judd & Irish, 2009; Kilgore et al.,
1997; Smith & Jones, 1910) that provide a record for conflicts within Nubia (Judd, 2006) and
with colonizers from Egypt (Buzon & Richman, 2007). Comparison of rural with urban
individuals reveals a distinctive difference between the two, suggesting a difference in risk
and lifestyle. That is, urban individuals have more cranial injuries, likely reflecting the
availability and use of hand-held implements, such as walking sticks. However, aside from
gross comparisons of fracture prevalence, there is surprisingly little information on temporal
trends in violent trauma, whereas temporal comparisons reveal important trends in accidental
injury patterns in recent humans (also see later).
The earlier record from the regions shows a highly elevated prevalence of injury. At Semna
South (c. 2000 BC) and Kerma (c. 1750–1500 BC), postcranial fracture prevalences exceed 20%
(Alvrus, 1999; Judd, 2002). Many of the individuals in these settings had multiple injuries,
mostly arising from accidental causes (Judd, 2002). For example, in a rural Kerma culture
sample (2500–1550 BC), of 55 adults, 80% had an injury, with 62% having at least two
injuries. This is higher than the urban sample at Kerma, strongly suggesting that lifestyle and
risk of trauma due to either occupational or environment circumstances was higher in rural
settings (Judd, 2006). Of those individuals with injuries, 51% of females and 71% of males had
multiple injuries.
The remarkably high level of endemic violence in Nubia was also fueled by external forces,
including repeated attacks from Egypt, due largely to the fact that Nubia controlled key trade
routes, which promoted a long history of raiding into Nubia. Kerma was able to resist these
forays for centuries, finally being defeated in c. 1520 BC (Buzon and Richman, 2007). While
warfare was an important policy of control, Egypt appears to have shifted its preference from
raiding and battle to diplomacy and incorporation of Kerman Nubians into Egyptian
government. Buzon and Richman (2007) tested the hypothesis that this shift in political
strategy would result in lower levels of violent confrontation and injury than what Judd
(2004) documented for earlier periods involving conflict between Nubia and Egypt. Their
analysis of remains from the Egyptian colonial settlement of Tombos (1550–1050 BC) in
northern Sudan revealed a remarkably lower prevalence of violence-related cranial injury in
comparison with the earlier Kerma and later Tombos populations, decreasing from 11.2% to
1.4%. Analysis of social and political circumstances strongly suggests that this change in
violence-related injury was due to alterations in Egyptian policies in Nubia.
In the Wadi Halfa region of Sudanese Nubia, fracture prevalence increased in the Christian
period (AD 550–1300) relative to earlier periods (Armelagos, 1969; Armelagos et al., 1981).
Within the Christian period in Kulubnarti, Nubia, there is a general increase in fracture
prevalence from 18% to 23% in the early (AD 550–750) to late (AD 750–1550) periods,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 157

respectively (Burrell et al., 1986; Kilgore et al., 1997; Van Gerven et al., 1995). Upper limb
fractures show an especially pronounced increase (30%) (Van Gerven et al., 1995).
Increase in skeletal trauma in juveniles and old adults is more pronounced than in other age
groups in the late Christian period at Kulubnarti. These age-specific increases in fracture
prevalence may be due to elevated risks of living in two-story houses in the late period versus
one-story houses in the early period (Burrell et al., 1986). Access to the living area on the
second story in late-period houses was gained by use of a retractable ladder, which may have
caused falls and other types of accidents (Burrell et al., 1986).
A significant portion of the Kulubnarti individuals possesses multiple injuries (27%). This
unusually high prevalence of accidental injuries, coupled with the presence of numerous and
severe fractures, reflects the hazards of living in a very difficult terrain. Unlike Lower Nubia to
the north, cultivated areas in Upper Nubia are highly constricted and are limited to small
pockets of flat land immediately adjacent to the Nile River. Individuals living at Kulubnarti
would have been exposed to difficult walking conditions on a daily basis. The adoption of
defensible architecture (e.g., two-story houses) later in the Christian period may also have
placed individuals at increased risk of injury.
Most Kulubnarti fractures are in the forearm (75% of fractures) (Kilgore et al., 1997).
Aggressive interactions may have contributed to some of the fractures because a push and
subsequent fall could have resulted in forearm fractures. The record suggests that the difficult
terrain and nearly continuous conflict between Egypt and Nubia resulted at various times in a
high risk of injury and death, arising from accidents and interpersonal conflict.

4.3.10 Australian foragers


Cranial trauma in prehistoric native Australian populations provides a compelling picture of
violence and injury in a wide range of geographical, ecological, and cultural settings (Webb,
1989, 1995). Most regions of prehistoric Australia have relatively elevated frequencies of
cranial trauma, especially depressed fractures (Table 4.2). Most of the injuries are well healed,
indicating that the attacker’s intentions may have been to injure rather than kill the victim.
Many studies of human populations worldwide document the higher proportion of violence
directed at males (Gurdjian, 1973; Jiménez-Brobeil et al., 2009; Lahren & Berryman, 1984;
Novak, 2006, 2014; Owsley, 1994; Robb, 1997b; Walker, 1989; and many others; but see
Wilkinson & Van Wagenen, 1993), which likely reflects the central role of men in the violent
resolution of conflicts in most human societies. This sex-specific pattern of head injury
contrasts with that of Australia. In virtually all samples throughout the continent, adult
females show a consistently higher prevalence of cranial injury than adult males, thus
contributing to the greater prevalence in females than males overall (Table 4.2). Some of
the sex differences in specific localities are slight, but many series show a remarkably strong
disparity between sexes. For example, in the south coastal Swansport sample, 39.6% (21/53)
and 19.3% (11/57) of females and males exhibit cranial trauma, respectively. For the few
skeletal series (4/22) where males have more cranial injuries than females, the differences are
statistically indistinguishable. The disparity in cranial injury between males and females is not
restricted to prevalence alone. For virtually all regions of Australia – regardless of ecological
or cultural setting – women show a predominance of depressed fractures on right parietal and
occipital bones. This pattern suggests that attacks came from behind the victim, perhaps while

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
158 Injury and violence

Table 4.2 Cranial trauma by region in Australia. Percentages of crania are shown (Adapted
from Webb, 1995: Table 8–2)
N One lesion (%) Two lesions (%) Three lesions (%)
Males
Central Murray (247) 13.4 3.6 0
Rufus River (122) 26.2 7.4 0.8
South Coast (138) 14.5 4.4 0.7
Desert (132) 16.7 3.8 1.5
Tropics (92) 6.6 0 0
East Coast (133) 23.3 6.8 1.5
Females
Central Murray (151) 19.9 4.0 1.3
Rufus River (83) 27.7 8.4 2.4
South Coast (123) 31.7 10.6 4.9
Desert (51) 33.3 11.8 5.9
Tropics (62) 24.2 9.7 4.8
East Coast (86) 32.6 10.5 3.5

fleeing the attacker. Adult males show a different pattern of injury location to that of women:
for the entire series, more left than right parietal bones are fractured. This pattern suggests
that male violence usually involved face-to-face confrontations.
It is not uncommon for an individual in the Australian samples to have two, three, or even
four cranial bones that display depressed fractures (Table 4.2). Consistent with the sex
differences in the prevalence of crania affected, more women than men have multiple injuries.
The general pattern, therefore, indicates that violence and aggression were directed more at
women than men in prehistoric foraging societies throughout Australia. Overall, however, the
pattern of higher injury in females than males evokes the conclusion that perhaps the record
represents continental-wide domestic violence – females being injured by male associates or
partners. Indeed, in modern societies, millions of women globally are injured by males in
domestic contexts (Novak, 2006).
In Australia, violence was not limited to prehistoric societies. Ethnographers observe that
violence is a common occurrence and a part of everyday discourse (Burbank, 1994). Unlike
Western societies, such as in the United States where fighting – and especially aggression against
women – is viewed as a deviant behavior, physical aggression in native Australians is considered
an accepted if not legitimate form of social interaction (Burbank, 1994; Myers, 1986). Burbank
(1994) provided detailed observations on physical aggression in men and women in an aborigi-
nal group living in Arnhem Land (Northern Territory). Her study shows that both men and
women were heavily involved in confrontations. However, the majority of aggressors and their
victims are adult females. These observations of both deceased and living native Australians
reveal a striking consistency in behavior between prehistoric and contemporary populations.
Women played a key role in aggressive encounters, and not simply as victims of attack. This
work underscores the important notion that violence, domestic or other, is not the sole purview
of males, certainly in this setting and likely others around the world (Guliaev, 2003).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 159

4.3.11 Southeast Asia: Iron Age Cambodia


The record of bioarchaeology in Southeast Asia reveals little evidence of violence and
associated trauma. However, new analyses of the Iron Age (350 BC–AD 200) Phum Snay
cemetery in northwest Cambodia reveals increased violence-related injury and competition
for arable land and other resources, in many ways providing a context for the later emergence
of complex societies and state systems involved in widespread conflict, such as Angkor after
AD 800 (Domett et al., 2011). In this setting, 30 of 128 individuals display cranial trauma, with
most in adult males (66%). The injuries comprise blunt-force (depression fractures) and blade-
induced trauma. Some of the trauma is perimortem, clearly resulting in death, whereas some is
healed. A number of individuals have multiple injuries of varying degrees of healing. None of
the postcranial regions show violence-related injuries. Thus, in this setting, combatants
focused their attacks on the heads of opponents. Like other series described previously (e.g.,
Norris Farms), the series shows that violence occurred over a period of time. That is, some of
the victims have clear evidence of previous confrontations, which they survived. Unlike Norris
Farms, however, the victims are predominantly adult males. Moreover, a number of the males
were associated with military paraphernalia, such as weapons in their graves. While the
cemetery certainly includes victims, it also has a focus on militarism and purposeful arming.
The placement of this series in this time and place supports the emerging understanding that
armed conflict arising from competition and social friction arose in the region centuries
before the appearance of complex social systems in Southeast Asia several centuries later.

4.3.12 Northern European violence


There is a rich bioarchaeological record for violence and confrontation in Europe for espe-
cially post-Pleistocene western and central Europe. Like the settings described earlier, the
social, political, and economic contexts are almost always related to competition for resources
generally, but local circumstances may also feature prominently.

Scandinavia and western Europe


The historical record of violence and warfare is abundant for northern and western Europe.
Systematic studies of violence have been produced for several areas of northern Europe,
including Scandinavia, and especially prehistoric and early historic Denmark. Analysis of
human remains reveals evidence of traumatic injury, decapitation, and mutilation. Like much
of the history of paleopathology, these studies are largely descriptive, having focused on
single or few individuals (Bennike, 1991a, 1991b).
Relying primarily on remains dating from the Mesolithic (c. 8300–4200 BC) to the Middle
Ages (to AD 1536), Bennike (1985) identified patterns of injury in Denmark. These patterns
can be characterized as involving a predominance of cranial trauma in mostly adult males
(principally depressed fractures) on anterior cranial vaults, indicating face-to-face violent
interactions. Folklore and historical accounts emphasize a high prevalence of violence during
the Viking period (AD 800–1050). Bennike’s (1985) assessment clearly indicates that the
Mesolithic and Neolithic periods were far more violent than the Viking period: Mesolithic
crania display the highest prevalence of trauma (43.8%), which is markedly reduced in the
Neolithic (9.4%), Iron Age (4.7%), Viking period (4.3%), and Middle Ages (5.1%). Violence is

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
160 Injury and violence

well illustrated by the presence of projectile injuries, sword and axe cuts, cranial depressed
fractures, and decapitation (Bennike, 1985; Ebbesen, 1993; Kannegaard Nielsen & Brinch
Petersen, 1993). At the Mesolithic site of Bogebakken, a bone projectile was found lodged
between the second and third thoracic vertebrae of an individual. In fact, all projectile wounds
at this and many other Danish sites are found in the thoracic and head regions, revealing the
lethal intentions of the attackers.
Placement of Denmark in the larger regional context of western and central Europe reveals
that the Neolithic and the foraging-to-farming transition was a relatively turbulent period in
comparison with later times, and not only in Denmark (Fibiger et al., 2013; Smith, 2014;
various in Schulting & Fibiger, 2012). The reasons for violence in this broad expanse are
varied. For example, the co-occurrence of elevated violence-related injuries in the early
Neolithic (c. 5000 BC) and deteriorating environmental circumstances across the region
suggests the possibility of negative social, cultural, and economic factors (compare with
Orschiedt & Haidle, 2012; Teschler-Nicola, 2012; Wahl & Trautmann, 2012).
At least one Neolithic series from Talheim, Germany, represents a massacre (Wahl &
Trautmann, 2012). The posterior position of traumatic injuries for many of the 34 individuals
documented indicates that they were attacked from behind while fleeing from aggressors.
With a few exceptions, most cranial injuries (e.g., depressed fractures) throughout the region
are healed, suggesting that as in southern California coastal foragers, inflicted injuries were
not intended to have a lethal outcome (Walker, 1989; and authors cited earlier). In their
analysis of cranial trauma in Neolithic Denmark, males generally have more anterior and left-
sided trauma than females, indicating the involvement of men in face-to-face encounters
(Fibiger et al., 2013).
In interments dating to the Middle Ages, the heads of victims had been removed and placed
between their legs. The reasons for this unusual treatment are unclear, but during the Middle
Ages, the practice was associated with criminals in order to prevent their return following
death (Bennike, 1985; and see later). Decapitation and other forms of head and neck trauma
were likely more common than is indicated by the skeletal evidence alone. A number of
Neolithic and Iron Age bog corpses show evidence of decapitation and strangulation, the
latter of which may have involved hanging (Bennike, 1985; Glob, 1971). Owing to the
relatively small number of projectile- and weapon-related deaths, it is not possible to identify
a pattern of decrease or increase in violence-related death in Denmark (Bennike, 1985).
However, there is a shift from the use of projectile weapons to axes and swords in the Iron
Age (Bennike, 1985). The lethal nature of this new weaponry for the enemies of Danes is
demonstrated in at least one battle site (see later).
Violence in western Europe during the Mesolithic and Neolithic was likely highly localized,
with a relatively higher prevalence in some regions than in others (Bennike, 1985; Frayer,
1997; Jiménez-Brobeil et al., 2009; Orschiedt et al., 2003; Papathanasiou et al., 2000; various
in Schulting & Fibiger, 2012). For western Europe as a whole, violent trauma was frequent
during the Neolithic and Mesolithic (Schulting & Fibiger, 2012; but see Constandse-
Westermann & Newell, 1984; Roksandic et al., 2006; Schulting, 2006). An increase in
population density, increase in social complexity, and resource circumscription during the
period suggests the potential for an increase in hostilities. Moreover, there are clear instances
of violence predating the Neolithic, indicating possible processes underlying violence in the
region; for example, at the Mesolithic site of Ofnet (Germany) where two pits contain the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 161

skulls of 32 individuals. The skulls show perimortem depressed fractures and cutmarks. Some
of the skulls have upper cervical vertebrae articulated, a number of which display cutmarks
suggesting slitting of the throat and decapitation.

Battle of Wisby
The Middle Ages in Europe involved a tremendous upsurge in armed conflict between many
warring city-states and various confederations of states. Today, across Europe, the vestiges of
walls surrounding villages and towns remain, a lasting record of violence and threat of
violence during this time. Only rarely, however, have battle sites from the conflicts resulted
in documented archaeological remains. One important exception is the battle site of Wisby,
located on the island of Gotland in the Baltic Sea. Hundreds of skeletons excavated at the
battle site present some of the grim details of preindustrial warfare in northern Europe.
The city of Wisby was invaded in 1361 by an army led by the Danish king, Waldemar
(Thordeman, 1939). Over the course of a single day, the poorly organized peasant forces
defending the city were decisively defeated and massacred by the king’s highly disciplined
army. Estimates indicate that some 1800 Gotlanders were killed in this battle (Ingelmark,
1939). Archaeological excavations of three common graves at Wisby yielded an enormous
sample of human skeletal remains (n¼1185). Analysis of these remains reveals that only
males were victims, but the age distribution was extraordinarily varied, ranging from adoles-
cents to very old adult males (Ingelmark, 1939).
Consistent with research completed on skeletons from other Middle Ages archaeological
sites (compare with Bennike, 1985), most of the injuries resulted from the use of cutting
weaponry, especially swords and axes (n¼456). A significant minority of injuries was from
projectiles (n¼126). Skeletal wounds are variable, ranging from scratches and nicks to
dismemberment. The latter, for example, is illustrated by the presence of severed hands and
feet, partial limbs, and complete limbs. One individual expresses the intensity of battle: the
lower halves of both left and right tibiae and fibulae are completely severed. The lower legs are
affected more than any other area of the body: about two-thirds (65%) of cutting trauma
involved tibiae. Ingelmark (1939) observed that the focus on the lower limbs likely reflects the
use of shields and protective clothing for the body trunk, leaving the legs especially vulner-
able to injury. Sword blows directed at the lower legs typically resulted in the slicing and
chipping of bone on the anterior crests of tibiae.
Poor protection of heads of individuals from the Gotlander army is indicated by the large
number of cranial injuries, some of which involve extremely deep cuts. Heads of some groups
of Gotlander soldiers may have been better protected than others. For example, only 5.4% of
crania are injured in common grave no. 3; this frequency contrasts with the prevalence of
cranial injuries of 42.3% and 52.3% in common graves no. 1 and no. 2, respectively. The
majority of cranial wounds are on the left side of the head, which fits the expected pattern of
injury sustained by a weapon held by a right-handed individual during a face-to-face
encounter (Courville, 1965). Some crania have injuries on posterior vaults, suggesting that
these victims were struck from behind while fleeing their attackers.
The presence of all ages of males suggests that the majority of the male population in and
around Wisby were recruited for the defense of the city. Analysis of pathological conditions
suggests that virtually anyone who could walk (and even those who could not) were drafted.
Ingelmark (1939:192) remarked on the “good many morbid processes” present in the skeletal

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
162 Injury and violence

assemblage of battle victims. Many vertebrae have pronounced osteoarthritis, and at least
four individuals have extensive vertebral tuberculosis. One individual displays a completely
ankylosed (fused) knee. The angle of flexion (about 55) greatly disabled the individual:
running was impossible for this victim. A number of individuals show well-healed but poorly
aligned femoral neck fractures, which limited their ambulatory capabilities. These observa-
tions, combined with other health problems – including skeletal infections and numerous
healed, but malaligned limb fractures (n¼39) – also contributed to reduction in efficiency on
the battlefield. The defending army, therefore, was not composed of a group of robust males
who were in their peak years of fighting prowess. Like many of the skeletal samples discussed
in this chapter from both New and Old World contexts, these victims of the massacre were
members of a population not unfamiliar with violence during their lifetimes (compare with
Milner et al., 1991; Willey, 1990).
The pattern of injury in the Wisby series is similar in many respects to the pattern of injury
documented in more detail from the Towton battle site dating to AD 1461, a century after
Wisby. There, some 28 000 of an estimated 100 000 combatants died on a single day (Fiorato
et al., 2000; Knüsel, 2014). Analysis of 37 of the casualties from a mass grave revealed clear
indicators of perimortem trauma, ranging from single to multiple (up to nine) injuries (Novak,
2000). Interestingly, nine of the 29 crania examined had well-healed trauma, likely derived
from previous battle-related injuries. Most of the injuries were on the left side of the face and
vault, indicating blows struck by a right-handed person successfully landing the business end
of the weapon on the victim in a face-to-face confrontation. The injuries include sharp (blade),
blunt, and projectile wounds (Figure 4.19). The injuries provide a picture that contrasts with the
romanticized version of Medieval battle showing honor and pageantry, but show clearly that
those who had inflicted the injuries were well trained in the combat methods and practice. The
patterns provide a useful record for evaluating the consequences of conflict in this key setting
prior to the use of firearms in modern warfare (Cunha & Silva, 1997; Düring, 1997; Giuffra,
Baricco et al., 2013; Mitchell et al., 2006; Šlaus et al., 2010, 2012; Weber and Czarnetzki, 2001).

Beheading in Britain
In the upper Thames valley, a high frequency of decapitation and prone burial in Romano-
British (third to early fifth centuries AD) cemeteries (Cassington, Radley, Stanton Harcourt,
Queensford Mill, and Curbridge) has been documented (Harman et al., 1981). In total, 15.3%
show evidence of beheading. Analysis of other Romano-British cemeteries indicates that this
practice was part of a widespread behavior during this period (Anderson, 2001; Bush & Stirland,
1988; Montgomery et al., 2011; Philpott, 1991; Tucker, 2014; Wells, 1982). During the
following Anglo-Saxon period (fifth to tenth centuries AD) beheadings continued, and victims
were included in execution cemeteries (Buckberry, 2008). Reminiscent of the decapitations from
Denmark (Bennike, 1985; and see earlier), heads of a number of burials had been purposefully
placed between and associated with the legs of the deceased (Anderson, 2001; McKinley, 1993).
The simultaneous occurrence of decapitation and prone burial in Romano-British and Anglo-
Saxon cemeteries suggests a probable connection between the two practices. The demographic
composition of decapitated and prone skeletons shows a selectivity for adults, suggesting that
execution may have been the primary motive. Review of historical, archaeological, and folklore
literature indicates other possibilities, such as prevention of the deceased from walking or
communicating, sacrifice, and deprivation of the soul, for either sacrificial purposes or for

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 163

Figure 4.19 Sharp force trauma to the face associated with a bladed weapon; Towton, United
Kingdom. (From Novak, 2000; reproduced with permission of author and Oxbow Books.)
(A black and white version of this figure will appear in some formats. For the color version,
please refer to the plate section.)

punishment for some wrongdoing (Harman et al., 1981). Decapitation and/or prone burial, may
have been a “final form of indignity inflicted on the corpse of an individual in consequence of
particular characteristics or offenses during life. But it seems more probable that both were
believed to have some effect on the subject in an after life” (Harman et al., 1981:168).
The manner of beheading is indicated by the location and pattern of cutmarks on affected
skeletal elements. Severing of the head was usually done in the upper neck region. Damage to
anterior surfaces of cervical vertebrae in some individuals and posterior surfaces of cervical
vertebrae in others indicates that the beheading blow was delivered from in front and behind
at various times, and probably with a variety of tools (Harman et al., 1981; and see McKinley,
1993; McWhirr et al., 1982). Detailed analysis of a beheading victim from Hertfordshire shows
a series of at least six carefully placed cuts, including three cuts on the anterior odontoid
process, superior body, lower body, and right inferior articular process (McKinley, 1993). The
narrowness of the cuts indicates that the decapitation was completed with a narrow blade
administered as blows to the neck.

4.3.13 European invasion of the Americas


The patterns of violence and warfare discussed previously for Europe set the context for
understanding the bioarchaeological record of violence in the postcontact, colonial-era
Americas. When Europeans began exploration of the New World in the late fifteenth century,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
164 Injury and violence

they brought with them a weapons technology that facilitated their (mostly rapid) conquest of
native populations. The early expeditions were violent affairs at times, resulting in brutal
treatment of natives (Weber, 1992). Although these tactics seem repulsive now, they were well
within the bounds of behavior regarded as fully acceptable for European males during the
Middle Ages. Historical literature and accounts of violent confrontations (e.g., Wisby) indicate
that conflict behavior between European males was often excessively cruel (Weber, 1992). The
study of native remains dating to the early period of contact with Europeans has provided a
new dimension to understanding the nature of the interactions between these groups.

La Florida
The study of hundreds of human skeletal remains from colonial-era Spanish sites in the
American Southeast (Georgia, Florida) and Southwest (New Mexico and Arizona) provides
important perspective on violent confrontation, especially during the sixteenth and seven-
teenth centuries. In Spanish Florida (present northern Florida and coastal Georgia), the region
named La Florida by Juan Ponce de León in 1513, short-term encounters between native
populations and Spaniards occurred during the exploration period (c. 1513–1565), followed
by long-term, sustained contact during the mission period (1565–1704) (McEwan, 1993). The
earliest contacts frequently resulted in hostile interactions and deaths of both Europeans and
natives (Varner & Varner, 1980). The later mission period was relatively peaceful; long periods
of calm were occasionally punctuated by native revolts violently put down by Spanish
military forces (Hann, 1988).
Analysis of skeletal remains from both periods of Spanish occupation in the region
produced only limited evidence of violent interactions. In skeletons from the Tatham Mound
site on the Gulf coast of western Florida – the probable location of Hernando de Soto’s visit
in 1539 – perimortem trauma caused by the impact of metal weapons is present in 17 skeletal
elements (Hutchinson, 1991, 1996, 2006). The most dramatic examples of cut bones include
the severed acromion process of a right scapula and a left humerus diaphysis cut through
60% of the midshaft (Figure 4.20). Neither bone showed evidence of healing. Other long
bones in the sample have multiple cutting injuries around diaphyseal perimeters. In total, the
pattern of damage appears to be associated with purposeful dismemberment (Hutchinson,
1996). It is possible that Indians inflicted the injuries using captured Spanish weapons. The
early dates of the site (early sixteenth century) suggest that this is an unlikely possibility.
The injuries were more likely inflicted by Spaniards (Hutchinson, 1996). The pattern of
injury is also consistent with European battle tactics. For example, the orientation in the
scapular cut is consistent with the practice of removing the fighting arm of the opponent
(Hutchinson, 2006).
Only one other skeleton in Spanish Florida shows evidence of violent confrontation.
A single high-status male from Mission San Luis de Talimali (AD 1656–1704) likely died
from a gunshot wound, but it is not possible to identify the perpetrator, Indian or Spaniard
(Larsen, Huynh et al., 1996). Thus, based on the study of violent trauma in skeletal remains,
the legends suggesting an unusually cruel treatment of natives – at least as it is indicated by
metal-edged or firearms weaponry – are not substantiated. Rather, the form of violence that
has been well documented in this setting is what Klaus (2014a) described as bioarchaeological
evidence of structural violence (see later).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 165

(a)

(b)

Figure 4.20 Cut adult humerus (a) and scapula (b); Tatham Mound, Florida.
(From Hutchinson & Norr, 1994; reproduced with permission of authors and
John Wiley & Sons, Inc.) (A black and white version of these figures will appear in some
formats. For the color version, please refer to the plate section.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
166 Injury and violence

American Southwest
There is a highly visible record of violence in precontact Southwestern populations, much of
which highlights episodes of stress and (possibly motivated by starvation and heightened
resource stress generally) cannibalism. This record is closely linked with periods of violence as
shown by an abundant skeletal record of injury and violent trauma, sometimes in association
with processed remains (Turner & Turner, 1999; White, 1992). The presence of fortifications
and other defensive architecture and high frequencies of traumatic injury (including injury
resulting from violence) in some later prehistoric sites suggests that confrontations were
frequent (Stodder & Martin, 1992). Cranial trauma in the American Southwest shows an
increase in prevalence during the late prehistoric period, which continued into the historic,
mission period (Bruwelheide et al., 2010; Stodder, 1990, 1994; Stodder & Martin, 1992).
Archaeological and historical evidence shows that the high frequency of cranial trauma in
the historic period can be attributed to confrontation between Spaniards and Indians as well
as among Pueblos, and between Pueblos and non-Pueblo native groups. The study of skeletal
remains reveals that most cranial injuries are in adult males (Stodder, 1994). At San Cristobal
and Hawikku, sites with significant contact-period skeletal assemblages, very high frequen-
cies of cranial injuries have been reported (Stodder, 1994). Twenty percent and 17% of males
have cranial trauma from the San Cristobal and Hawikku sites, respectively. Paleopathological
markers of stress (e.g., dental defects) indicate that nutritional and other health disruptions
were generally widespread during the late prehistoric and contact periods, which may have
contributed to fostering intra- and inter-group hostilities during this time (Stodder, 1994;
Stodder et al., 2002). At Pecos Pueblo, the prevalence of cranial trauma, largely in the form of
depressed fractures, in the mission population (42.1%) is considerably higher than the
prehistoric ancestral population and most other samples in North America where data have
been collected systematically (Bruwelheide et al., 2010).

4.3.14 North American military campaigns


The fight for political domination of vast areas of North America, especially after the
seventeenth century, is indicated by the many military campaigns involving confrontations
between warring European nations prior to American independence, between the fledgling
United States and United Kingdom, and between native populations and various Euro-
american or European interests.

Fort William Henry


During the North American French and Indian War (called the Seven Years War in Europe),
France and Great Britain fought over control of the vast territories of the Northeast United
States and Canada. During the summer of 1757, the British garrison surrendered Fort William
Henry at the southern end of Lake George, New York, to French and Canadian troops and their
Native American allies (Starbuck, 1993). As part of the conditions of surrender, British soldiers
and dependents were allowed to leave the fort and return to British-controlled territory under
French protection. The French-allied Native American warriors felt slighted that scalps and
other prizes of warfare would not be forthcoming. In retaliation, the French-allied Indians
killed the remaining British troops at the fort. Warriors then proceeded to kill or capture the
hundreds of civilians and soldiers under the care of the French troops.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.3 Intentional injury and interpersonal violence 167

Analysis of the remains of five adult males buried in a mass grave within the fort indicate
clear evidence of violence-related injuries, bearing testimony to the historical and fictional-
ized (Cooper, 1919) accounts of the battle (Liston & Baker, 1996). Four of the five show
perimortem trauma to the legs reflecting injuries received during the siege of the fort. None of
the injuries are healed, suggesting that they died prior to or during the siege. The trauma was
not lethal, but serious enough so as to prevent their departure from the fort. These skeletons
display a range of perimortem trauma that likely represents injuries resulting in death. One
individual shows a series of four cutmarks on the posterior surface of the odontoid process of
the second cervical vertebra. Another individual expresses a series of radiating fractures
through the face and frontal bone, indicating crushing of the skull with a blunt object. The
pattern of trauma suggests that the soldier had been beheaded from behind. All five individ-
uals show notches, slices, and gashes in skeletal elements of the anterior and posterior trunk
(e.g., scapula, ribs) and pubic region. The morphology of cutmarks evinces the use of both
knives and axes in the mutilation of victims.

Snake Hill
Some of the most intense fighting between the British and Americans during the War of
1812 took place in the frontier region between Lake Ontario and Lake Erie along the Niagara
River (Whitehorne, 1991). During the four-month period in 1814 when American troops
successfully captured and held Fort Erie on the Canadian side of the river, heavy siege and
combat resulted in the deaths of hundreds of soldiers from both the British and American
armies. Archaeological excavations at the battle site of Snake Hill, Ontario, resulted in the
recovery of the skeletal remains of American soldiers from burial and medical waste pits
(Owsley et al., 1991; Thomas & Williamson, 1991). Demographic assessment of the complete
or nearly complete skeletons indicates that they were young adult males, aged 15 to 29 years;
seven soldiers were older than 30 years at death. Half (50%) of the individuals in the sample
had fractures caused by damage from firearms projectiles. The general lack of healing in most
cases indicates that these wounds were usually fatal. The highest percentage of fractures were
ribs (28%; 7/25), followed by femora (25%; 7/28), and crania (9.1%; 2/22).
Locational assessment of skeletal wounds indicates that most injuries (69.8%) were above
the waist. With regards to the total number of noncranial and nonvertebral trauma (n¼53),
twice the number of fractures occurred on the left side (54.7%; 30/53) than on the right side
(26.4%; 14/53) of battle victims. This pattern may reflect handedness or body postures during
the battle (Owsley et al., 1991). Cause of death is especially apparent for several victims. For
example, a young adult died of a massive head injury whereby a firearm’s projectile had
passed through the left and then right parietal bones. This individual also had a large,
completely healed cranial depressed fracture from earlier injury. Other individuals had frac-
tured facial bones from a firearms projectile and shattered long bones.

Battle of the Little Bighorn


In present-day South Dakota in June of 1876, General George Armstrong Custer and 267 sol-
diers and civilians were overwhelmed and massacred by a superior force of Native Americans
(Scott et al., 2002). Reminiscent of prehistoric and historic conflicts between native groups in
the region (compare with Crow Creek, Larson Village site; and those discussed earlier), this

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
168 Injury and violence

battle is part of an overall pattern of political domination and control of lands and resources
in the Great Plains by opposing groups.
Within two days of the battle, eyewitness accounts described mutilation (including
scalping) and dismemberment not unlike patterns observed in other Plains samples (e.g.,
Crow Creek, Larson Village site; see earlier) (Scott et al., 2002; Snow & Fitzpatrick, 1989).
Temporary graves were hastily prepared at the locations where individuals were killed. Some
of the bodies were identified, but owing to decomposition and mutilation, many were not.
Skeletal fragments of battle victims from erosion and limited test excavations provide the
basis for detailed study of battle injuries (Snow & Fitzpatrick, 1989). Analysis of 375 partial
and complete bones and 36 teeth from a minimum of 34 individuals indicates three primary
types of perimortem trauma, including blunt-force trauma, cutmarks, and bullet wounds
(Scott et al., 2002; Snow & Fitzpatrick, 1989; Willey & Scott, 1996). Blunt-force trauma
involved catastrophic fragmentation of crania, and to a lesser extent, postcranial elements.
All 14 partial crania showed massive injuries due to heavy blows. Additionally, presence of
cutmarks on various skeletal elements indicates widespread perimortem mutilation. Several
different forms of cutmarks, ranging from fine cuts to pronounced incisions, reflect the use of
metal arrows or knives.
Cutmarks on a variety of skeletal elements indicate the high degree of mutilation of battle
victims. One individual, for example, has cutmarks on a humerus head and sternum. The use
of heavy metal-edged weapons (e.g., hatchets) is clearly indicated in several instances.
Elsewhere, a completely transected cervical vertebra indicates decapitation by a single blow
to the neck. Another individual shows distinctive sets of chopping blows to the proximal ends
of the left and right femora indicating purposeful dismemberment.
In addition to traditional native weaponry, the presence of gunshot wounds in six individ-
uals indicates the use of firearms by Native Americans at the battle site (Scott et al., 2002;
Willey & Scott, 1996). Individual M85, for example, had at least two upper body gunshot
wounds, including an entrance wound on a rib margin and shattered ribs from another
wound. A third wound is represented by a bullet or bullet fragment embedded in the distal
left radius. This individual also displays cutmarks on his clavicle. Three gunshot wounds are
located in the crania, one entering from the back and two entering from the right side and
exiting from the left.
In summary, based on the study of human remains from the battle site, a sequence of events
can be reconstructed: namely, soldiers were wounded, killed (frequently with blunt-force
trauma to the skull), and mutilated (Snow and Fitzpatrick, 1989). A consideration of direction
of entry wounds is consistent with historic records indicating that the battle was chaotic
(Willey & Scott, 1996). As would be expected, except for the use of firearms, the pattern of
killing and mutilation of victims is strikingly similar to that observed in other North American
native populations from the Great Plains and Midwest discussed in this chapter (e.g., Crow
Creek, Larson Village site, Norris Farms).

4.4 Medical care and surgical intervention


Depending on the severity of injury – originating from accidental or violent circumstances –
the survivor is often debilitated and unable to perform key functions, such as acquisition of
food and other essential resources. For purely economic reasons, it is in the best interest of the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.4 Medical care and surgical intervention 169

social group to ensure that the injured person returns to a functional state of health and well-
being. Ethnographic and historic accounts of nonindustrial societies report tremendous
variation in the care of traumatic injuries, ranging across alignment of fractures and use of
splints, immobilization, oral medicines, and other treatments (Ackerknecht, 1967; Ortner
2003; Roberts, 1991). For example, lack of angulation or significant difference in length of
long bones in the fractured versus the normal side has been documented in the Libben
population from the American Midwest (Lovejoy & Heiple, 1981) and in Medieval populations
from England (Grauer & Roberts, 1996).
Similarly, in prehistoric Australian skeletal series, most fractures show proper unification
and alignment (Kricun, 1994; Webb, 1989, 1995), which Webb regards as evidence for “a firm
commitment to care and concern for the injured patient” (1995:200). The presence of well-
healed amputations in two individuals from the central Murray Valley region suggests
knowledge of this type of surgical procedure before the arrival of Europeans (Webb, 1995).
In five to six thousand Nubian skeletons, some 160 fractures are present, most of which are
well-healed and aligned, with little evidence of infection (Smith & Jones, 1910). Bark splints
were found associated with limb fractures in a couple of instances (Smith, 1908).
Some earlier societies appear to have lacked either the ability or interest in treatment of
fractures. For example, the fourteenth-century Wisby battle victims had a high degree of
angulation of healed fractures, suggesting only minimal levels of treatment (Ingelmark, 1939).
The battle victims are largely drawn from the peasant class, and may not have had access to
the same medical care afforded the nobility.
Temporal assessment of a large sample of Roman, Anglo-Saxon, and Medieval long bone
fractures reveals changes in injury management in these populations (Roberts, 1991). During
the earlier Roman and Anglo-Saxon periods, healing of fractures was generally good,
suggesting that fracture sites were correctly reduced and aligned, probably with some type
of a support (Roberts, 1991). Treatment was so widespread in the Roman period that
deformities from poorly healed or misaligned fractures were no more common than they
are in living populations. Medieval management of fractures appears to have been less
efficient than that in earlier periods. There is a generally higher prevalence of deformation
and angulation of long bones (Roberts, 1991). Additionally, many fractures (35/59) from this
period had associated periosteal infections. If this analysis and the findings based on the
study of the Wisby sample are representative, it appears that northern European populations
living during the Middle Ages were far less knowledgeable about fracture management than
their forebears.
Treatment of head injuries is suggested by the association between trephination and cranial
trauma in a number of settings. In Denmark, trephinations frequently accompany cranial
trauma, including fractures and sword or axe cuts (Bennike, 1985). Most of the Danish
trephinations are on the left side of the cranial vault, which coincides with the location of
skull injuries received in battle. This pattern is also consistent with the predominance of
trephinations in males, the primary participants in battle and interpersonal conflict. Similarly,
over 50% of recorded trephinations from Anatolia documented by Erdal and Erdal (2011) were
associated with cranial trauma.
Precontact sites in Andean South America also show an abundance of trephination,
especially in Peru and Bolivia where scores of cases are reported. Analysis of crania from
central and south coastal Peru and regions of the highlands of the Central Andes reveals that

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
170 Injury and violence

Figure 4.21 Incomplete trephination and depressed skull fracture on right frontal bone;
Cinco Cerros, Peru. (Verano, 2007; reproduced with permission of author and University of
Arizona Press.) (A black and white version of this figure will appear in some formats. For
the color version, please refer to the plate section.)

adult males comprise the majority of trephined individuals, but adult females and some
juveniles are also included (Verano, 2003). The association between cranial injuries and
trephination indicates that this form of surgery was likely performed as a treatment for head
trauma (Figure 4.21). Diachronic comparisons indicate that the frequency of well-healed
trephinations increased over a 2000-year period (400 BC–AD 1532). The highest rate of
success was from the latest precontact period (Inca), including some individuals with as many
as five to seven healed trephinations. The apparent increase in survival may have been due to
the reduction in size of the trephination opening as well as to the greater use of the circular
grooving technique of excision, thus reducing the risk of dura mater penetration and neuro-
logical damage (Verano, 2003).
The lack of an association between trephination and head injury in other settings suggests
that there may have been other motivations, including treatment of real or imagined ills. For
example, none of the few trephined crania from North America are associated with cranial
injury (Ortner, 2003; Stone & Miles, 1990).
Evidence for treatment of dental disease has been identified in the form of alterations of
teeth, namely drilling in tooth roots (Schwartz et al., 1995) and crowns (Bennike, 1985;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.4 Medical care and surgical intervention 171

Figure 4.22 Anteromedial (above) and scanning electron microscopic (below) perspective
of drilled right mandibular canine; Sky Aerie Overlook, Colorado. (White et al., 1997;
reproduced with permission of authors and John Wiley & Sons, Inc.)

Koritzer, 1968; Turner, 2004; White et al., 1997). These holes are usually found in association
with carious or otherwise diseased teeth, indicating a therapeutic intention (Figure 4.22).
In summary, the study of samples worldwide reveals that injuries sustained either by
accidental means or during violent confrontations were treated in some earlier societies.
These findings indicate that many past societies were aware that proper restoration of
function could only be brought about by appropriate treatment protocols (Roberts, 1991).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
172 Injury and violence

4.5 Interpreting skeletal trauma


A variety of injuries are well documented in earlier societies. Accidental injuries generally
reflect the hazards of day-to-day living, including during food procurement and preparation,
specific kinds of occupations, and in transportation from one place to another in general. The
inchoate analysis presented in this chapter indicates that walking on difficult terrains or
engaging in behaviors requiring levels of high activity tend to result in elevated prevalences
of skeletal injuries, such as lower limb, Colles’ and rib fractures.
An abundance of skeletal data exists on violence and trauma, much of which is now placed
within the context of society, economics, and living circumstances. Violence plays a critical
role in human social relations and interactions, and is a strategy for dealing with a range of
circumstances, including expansion of political control and gaining access to resources.
Bioarchaeology couches its broader discussions within a contextual record, thereby rendering
violence in the past a subject of broader interest to anthropologists and other social scientists,
in developing a more informed understanding of social relations and the measurement or
documentation of conflict and warfare and the circumstances under which they arise.
Studies emphasizing the integration of social, political, or economic systems as they affect
conflict-related behavior are beginning to emerge for a number of regions. For example, in the
American Midwest and Southeast, numerous cases of projectile injuries, scalping and other
forms of mutilation, and lethal trauma covering a long temporal span are reported, which
provide compelling evidence of violence and warfare in prehistoric societies (Milner, 1995).
This evidence suggests that there is violence relatively early in prehistory (e.g., western
Tennessee Valley; Smith, 1995, 1997), but temporal comparisons indicate that conflicts
leading to injury increased in frequency, especially during the late prehistoric period (post-
AD 1000) in the Eastern Woodlands of North America. This regional trend appears to be
related to an increase in social tensions due to population increase, sedentism, increasing
social complexity, and increased focus on restricted and valued resources, especially high-
yielding domesticated plant foods (Bamforth, 1994; Eisenberg, 1986; Milner, 1995; Milner
et al., 1991). This evidence runs counter to the arguments raised by various authors (Ferguson,
1990, 1995; Ferguson & Whitehead, 1992) who contend that violence is either missing or
minimal in precontact New World societies, having been a result of disequilibrium arising
from Western contact. Certainly, social and cultural disruptions arising from contacts with
expanding Western, state level societies resulted in increased conflicts within some regions
(Bamforth, 1994). The skeletal evidence from several well-studied regions indicates, however,
that conflicts leading to injury were commonplace in a variety of settings. Skeletal injury
resulting from violence, therefore, represents a principal indicator of environmental stress in
human populations.
It is important to recognize variability in violence-related injury within larger regions. For
example, in the Tombigbee Valley of Alabama, a decrease in skeletal injuries due to violence
coincided with an increase in dispersion of human settlement and increased political central-
ization from the period of about AD 900–1550 (Steponaitis, 1991). Thus, reduced circumscrip-
tion of the population, perhaps brought about by new political forces, likely had an influence
on conflict in this setting. The influence of political factors may explain why some regions
undergoing an increase in population density and social complexity do not show an increase
in violent trauma (e.g., Mesolithic Europe: Constandse-Westermann & Newell, 1984). Violent

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.5 Interpreting skeletal trauma 173

injury occurs in individuals and populations who were the victims of conflict. Thus, cemetery
assemblages representing groups who were the winners would not be expected to exhibit the
frequency of injury seen on the losing end of violent encounters.
Elevated prevalence of violence and injury mortality in some prehistoric and historic
settings and their possible relationship with increased population density and/or resource
circumscription is similar to that of recent, twentieth-century populations. For example,
Relethford and Mahoney (1991) documented markedly higher rates of injury mortality in
the most densely populated areas of New York State (excluding New York City). These
similarities may reflect common themes between past and recent humans, such as high
density of population and social inequalities that serve to promote violence. Population
density is a complex composite of a number of factors, such as the physical and sociocultural
environments, demographic, cultural, and social influences, and individual behavior. There-
fore, although these apparent similarities are informative, it is important to identify specific
causal factors in specific settings before making conclusions generally regarding the relation-
ship between population distribution and injury mortality in humans.
In some regions, clear patterns of violence have begun to emerge. For example, an increas-
ingly hostile landscape in the millennium preceding European contact in the Eastern Wood-
lands generally is corroborated by the archaeological evidence of an increase in defensive
construction in later prehistory (Keener, 1999; Milner et al., 1991). It is during this time that
there is the rise of chiefdoms, population increase, and competition between neighboring
villages. Similar patterns of increasing conflict and aggression in the final centuries before
European conquest are indicated in the southern California Santa Barbara Channel islands,
the American Southwest, Southeast, Great Plains, and Arctic (Lambert, 2007, 2014). Regional
investigations suggest that resource productivity and climatic instability may have had a
strong influence on the presence or degree of conflict in the past. This hypothesis is supported
by the increasing evidence for broad patterns of violence in association with climatic instabil-
ity and drought in the later prehistory of North America. In particular, some of the most severe
violence occurs in settings and periods where climate is unstable and prone to drought. This is
not to say that climate causes violence. Rather, periods of drought were triggering events,
setting off a stream of events resulting in violent encounters between groups competing for
the same limited resources. Similar patterns of the increase in evidence of violence are also
documented in a wide range of Old World settings (Kennedy, 1994; various in Knüsel & Smith,
2014; Schulting & Fibiger, 2012).
Overall, bioarchaeological studies indicate that violence and conflict are not random
events, but are strongly influenced by extrinsic factors, such as resource depletion and
competition for important resources. Dietary deprivation may have been a motivation for
cannibalism. Historical records in a number of settings are informative. For example, canni-
balism may have been symptomatic of a larger pattern of animosity and aggression between
groups (e.g., Saunaktuk) or a key part of ritualized violence involving sacrifice and
cannibalism (e.g., Teotihuacan).
The study of trauma in skeletal remains reveals that, on comparing different societies, areas
of the body attacked are highly patterned. Walker (1997, 2001a) observed that in modern
industrial Western societies (e.g., United States), the head and neck are highly favored targets
of attack, probably for both strategic and symbolic reasons. He argues that the face is an
appealing target because the injuries are especially painful. The face and head generally bleed

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
174 Injury and violence

profusely and bruise easily, which may symbolize the aggressor’s dominance (Walker, 2001a).
This probably explains why the most highly traumatized focal point of the body in recent
urban populations is the face (Allen et al., 2007; Hussain et al., 1994).
Many past societies show a penchant for head injury, but these injuries are usually directed
at the vault and not the face or dentition. Dental fractures are present in archaeological
remains, but they are relatively rare (Alexandersen, 1967; and see Leigh, 1929; Lukacs, 2007b;
Lukacs & Hemphill, 1990, for regional studies documenting dental trauma). The location of
the injury on the body provides insight into some of the details of conflict between the
individuals involved. For example, many cranial injuries are found on the left side of the
frontal bone or other anterior elements, indicating that a right-handed attacker successfully
engaged his/her weapon while facing the victim (e.g., native Australian males). A more
haphazard pattern of cranial injury (e.g., prehistoric Michigan) or higher frequency of trauma
on the right side or posterior vault indicate that injuries were sustained while the victim was
fleeing their attacker or perhaps while lying prone (e.g., Wisby). This pattern is more common
in women than in men in some settings (e.g., Australia, Michigan, Peru), suggesting that
aggression was also directed at women. Ethnographic evidence reveals that although the
aggressor was often an adult male, attack by adult females on other females (and on males)
occurs in no small number in some settings (Burbank, 1994).
Historical documentation indicates that children have long been a target of violent injury
and death. DeMause (1974), for example, regarded child abuse as widespread in Europe prior
to the eighteenth century. Yet, examination of thousands of archaeological skeletons reveals
remarkably little evidence of skeletal trauma – localized trauma-induced subperiosteal lesions
in multiple stages of healing, and perimortem fracture – associated with battered-child
syndrome (Fibiger, 2014; Lewis, 2014; Walker, 2001b; Walker et al., 1996; Wheeler et al.,
2013). Certainly, juveniles in earlier societies were victims of homicide and violence (e.g.,
Ofnet, Crow Creek, Norris Farms). Juvenile skeletons, however, lack the injuries associated
with long-term abuse. This suggests that, as with the pattern of facial and other injuries in
twentieth-century Western societies (Love et al., 2011), child abuse resulting in severe skeletal
trauma is primarily a modern phenomenon. Walker (2001b) suggests that the rise of childhood
abuse is due to the loss of social controls over behavior in largely recent urban settings in
comparison with controls present in earlier and traditional societies.
Technological factors are important in interpreting patterns and types of skeletal injuries.
The introduction of the bow-and-arrow is linked with an increase in lethal conflict (e.g.,
southern California). Prior to the invention of firearms, violence-related injuries were caused
primarily by projectiles, cutting, and blunt force.
The skeletal record shows the presence of both lethal and nonlethal forms of trauma, thus
providing essential insight into the previous history of interpersonal aggression both within
and between past societies. The Wisby, Towton, Crow Creek, and Norris Farms victims display
numerous healed injuries (e.g., cranial depressed fractures) that reflect a long and well-
established history of conflict well prior to the event or events resulting in widespread death.
In a sense, then, these injuries foreshadow a future act (e.g., a major battle) that later resulted
in more widespread violence.
The study of human remains from these sites suggests that debilitating injuries, poor health,
or generally high levels of physiological stress may have increased the susceptibility of a
population to attack and defeat. The Norris Farms and Crow Creek skeletons display numerous

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.5 Interpreting skeletal trauma 175

pathological indicators of stress, including iron deficiency anemia, dental defects, tubercu-
losis, and generalized infection, that reflect compromised health and the reduced ability to
perform subsistence and other arduous tasks (Bamforth, 1994; Milner et al., 1991). Although
this pattern of poor health does not explain the demise of the population, it suggests that they
may have had a reduced ability to successfully mitigate hostile social environments. Some
populations display fractures and other debilitating conditions that limited their ability to
protect themselves or even flee a more powerful adversary (e.g., Wisby).
Although both nonlethal and lethal forms of violent injury are highly prevalent in many of
the populations discussed in this chapter, the dominance of one category over the other
informs our understanding of the intentions of the attacker. For example, the higher preva-
lence of nonlethal than lethal injury in a number of settings – Australia, Santa Barbara
Channel, Peru, and Easter Island – indicates that injury was meant to maim and not kill the
victim. Death of the opponent was clearly the preferred outcome of attack in the Middle Ages
of Europe (e.g., Wisby), late prehistoric Great Plains and Midwest (e.g., Crow Creek, Norris
Farms), in the Arctic (e.g., Saunaktuk), and in historic North America (e.g., Snake Hill). In
prehistoric settings, it usually is not possible to determine the reasons for preference of lethal
or nonlethal forms of violence. In California, the shift to lethal forms of injury from projectiles
in later prehistory may have been influenced by the change in weapons technology coupled
with increasing resource stress.
Clear patterns of mutilation of victims are well documented in a number of prehistoric and
other New World settings. In North America, the evidence for removal of soft tissue and skin
from the cranial region – especially scalping – is abundant. Typically, the scalp was removed
by first slicing skin along the frontal and parietal bones and peeling back of the skin (e.g.,
Norris Farms, Koger’s Island), but other approaches involved removal of facial and other
tissues (e.g., Saunaktuk). Mutilation was a highly visible behavior. In addition to scalps,
tongues, noses, limbs, and heads were removed from the near-dead or deceased (e.g., Great
Plains). A more profound mutilation to the head than scalping was decapitation. Decapitation
was practiced in the New World and Old World, and for a variety of reasons. In Roman Britain
(and throughout history), this was a preferred form of execution in some groups. In northern
Europe, the head of the victim was sometimes placed between the legs (Denmark, England),
perhaps as the ultimate insult. Other unique and highly localized forms of body treatment of
the living victim were likely practiced. For example, gouging of knees at Saunaktuk in the
Arctic may have been associated with a practice of dragging the victim through the village
prior to his or her death.
Trauma data are important for dispelling prejudices and assumptions about past societies.
For example, hunter-gatherer societies around the world are often characterized as peaceful
inhabitants of stress-free environments living in a state of blissful repose (Lee & DeVore,
1968; Service, 1966; and see discussions by Burbank, 1994; Fienup-Riordan, 1994; Walker,
2001a). This characterization may reflect the fact that anthropologists doing fieldwork among
these societies are guests – after all, what guest is going to go back home and write about the
unpleasant things they observed, especially with regard to violent encounters between
individuals (Erchak, 1996; Keeley, 1996)? Many anthropologists underplay the negative or
offensive, avoiding realism in social beliefs. In fact, a number of cultures described as
nonviolent or peaceful have homicide rates far exceeding those of some Western nations
(Knauft, 1987).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
176 Injury and violence

Ethnographic research has undergone a dramatic change, with key developments showing
the social, political, and economic contexts for a variety of settings and the relevance of
violence to social life (Burbank, 1994; Kapferer, 2012; Stewart & Strathern, 2002). The point is
not to replace peaceful characterizations of earlier societies with violent ones. Rather, these
findings underscore the importance of substituting an incorrect image with one that fits the
evidence, past and present. This approach is critical for informing our perspective on past
groups as functioning societies rather than as images of what earlier social behavior must
have been like. This newfound precision adds a more complete historical context for the study
of recent human behavior. Anthropologists and others seem to employ – either consciously or
unconsciously – their own cultural and social assumptions about earlier societies in order to
“remember” the past (Keeley, 1996). We project these assumptions into a past that seems to
reflect current and highly biased perspectives on the condition of humankind, be they
peaceful or violent. These skeletal data help us to reconstruct and interpret trauma and
violence in a more comprehensive and accurate manner.
This chapter has discussed the obvious skeletal correlates of violence and trauma in a
wide range of societies around the world, namely injuries received in interpersonal
conflict, ritual, warfare, and other settings where individuals are intentionally injured or
killed. Yet, there is another kind of violence that places people at risk of harm or death,
perhaps not immediately, but rather the kind that inhibits a person from maintaining
homeostasis owing to social circumstances. This structural violence pertains to members of
societies where needs are not met or where they are exploited for economic, political, or
other reasons (Farmer, 1996). Borrowing from developments in other social sciences, Klaus
(2012, 2014a) makes the case that broken bones and cutmarks are not the only reflection
of violence in human remains. Specifically, he documents evidence of the impairment of
human needs caused by social structures that prevent persons from reaching biological,
ocial, and economic potentials in prehistoric societies involving social inequality. Focus-
sing on the colonial-era Lambayeque Valley, Peru, he draws on ethnohistoric and arch-
aeological context in the newly established Spanish colonial state. This hierarchical
structure contained strong elements of violent repression for members of the native
populations in the region. The dominant European power viewed native peoples as a
source ripe for exploitation and ruled by a set of repressive laws where their labor was
bought, sold, and (usually) abused. This kind of exploitative system denied certain foods
and other resources, yet at the same time, demanded heavy physical labor. The predictable
outcome, of course, would be evidence of poor health, disease, physiological stress, and
biomechanical demand. In these circumstances, this outcome in health is best described as
structural violence.
In order to test the hypothesis that structural forms of violence were in place in the
Lambayeque Valley, a region colonized by Spain beginning in 1534, Klaus (2014a) compared
the record of health and diet of precolonial populations with colonial-era populations
represented by remains recovered from the Chapel of San Pedro de Mórrope dating from the
period of 1536–1750. This record reveals a compelling picture of outcomes of repression,
including disrupted homeostasis and decline, a remarkable increase in porotic hyperostosis
(154% increase), enamel hypoplasia (184% increase), and periosteal reaction owing to infec-
tion (471% increase). Along with the bioarchaeological record of violence and traumatic
injury associated with the Spanish conquest elsewhere in the Inca Empire (Gaither & Murphy,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
4.6 Summary and conclusions 177

2012; Murphy et al., 2010), the record of structural violence provides a window onto a rapidly
changing political landscape involving conquest and population collapse.

4.6 Summary and conclusions


A diverse range of bioarchaeological evidence helps to inform our understanding of acciden-
tal trauma and violence-related injury and their relationship to behavior and lifestyle in
earlier societies. Human populations living in difficult circumstances have an elevated
prevalence of skeletal injuries due to accidental causes. The skeletal record of conflict in the
past is highly visible in a number of settings worldwide. The study of remains from a wide
variety of contexts helps to provide better understanding of the circumstances of violence,
whether due to intra- or inter-group conflicts. Some conflict may result in cannibalism, but
based on the study of human remains alone, it is difficult to identify causes of this practice.
Ritualized violence, including sacrifice and elaborate body treatment, is also highly visible in
some settings. Although limited in scope, its study provides an important link between culture
and treatment of the body both in life and in death. Regardless of the circumstances of death,
contextual data are essential for its interpretation, including resource availability, history of
intra- and inter-group social relationships, and weapons technology.
Many of the injuries we see in the archaeological past have clear analogs in the present,
including cranial blunt-force trauma, knife wounds, and projectile trauma of all kinds. An
analog that seems to be largely missing from the past is the battered-child syndrome. This
involves circumstances where a parent, parents, or guardians physically abuse, with violence,
young children under their care. The violence goes undetected for months, if not years,
resulting in skeletal lesions (Walker et al., 1997). The injury involving a distinctive pattern
of periosteal reaction is not present in archaeological remains (Walker, 2001b). This suggests
that some cultural norms of behavior that exist today were not present in the past.
The proximate circumstances for violence have been identified in a number of settings. For
example, in the American Southwest, the surge in violence and cannibalism coincides with
chronic drought. Historically, and in other archaeological settings, violence is often associated
with climatic instability that results in reduction of food from crop production. Today, many
world populations in developed nations live in circumstances where cultural buffering of
environment and food supply are not dramatically affected by climatic swings. This under-
scores the point that climate change and violence are not necessarily linked in a simple cause
and effect manner. On the whole, however, the record, past and present, documents a clear
causal link between climate and conflict (Hsiang et al., 2013).
The bioarchaeological record of violence and traumatic injury provides an important
dimension to our growing understanding of different levels of interpersonal interactions,
ranging from conflict between small groups competing for limited resources (e.g., Santa
Barbara Channel Islands, California; Norris Farms, Illinois; Talheim, Germany) to full-scale
subjugation, population expansion, and empire building (e.g., Wari, Inca, and Aztec empires).
The “success” of these endeavors is difficult to measure as both winners and losers display
evidence of traumatic injury. However, when viewed in context and drawing upon a range of
social, economic, and environmental evidence, the bioarchaeological record reveals key
findings that help us develop a more informed understanding of the origins and evolution
of violence and its various forms, both overt and structural.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:05:40, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.006
Activity patterns: 1. Articular
degenerative conditions and
5 musculoskeletal modifications

5.1 Introduction
Physical activity is a defining characteristic of human adaptations. Hunter-gatherers, for
example, are often characterized as highly mobile, hard-working, and physically active. In
contrast, agriculturalists are sometimes seen as having an easier life – they are settled in one
place and have a lighter workload than hunter-gatherers. In his popular and influential
archaeology textbook, Robert Braidwood (1967:113) distinguished hunter-gatherers as leading
“a savage’s existence, and a very tough one. . .following animals just to kill them to eat, or
moving from one berry patch to another (and) living just like an animal.” Ethnographic and
other research calls into question these simplistic portrayals of economic systems. Following
the publication of Lee and DeVore’s (1968) Man the Hunter conference volume, and especially
Lee’s (1979) provocative findings regarding work behavior and resource acquisition among the
!Kung in northern Botswana, a consensus emerged that, contrary to the traditional Hobbesian
depiction of hunter-gatherer lifeways as “nasty, brutish, and short,” prehistoric foragers were
not subject to overbearing amounts of work, and life overall for them was leisurely, plentiful,
and confident (Sahlins, 1972). More importantly, these developments fostered a wider discus-
sion by anthropologists and other social scientists of activity, behavior, and lifestyle in both
present and past hunter-gatherers (Kelly, 2013). These discussions led to the conclusion that
human adaptive systems are highly variable. As a result, it is now clear that it is not possible to
make blanket statements about the nature of workloads or other aspects of lifestyle in foragers
and farmers (Kelly, 1992, 2013; Larsen, 1995). Rather, workload and lifestyle are highly
influenced by the kinds of resources exploited, climate, and sometimes highly localized
circumstances. Nevertheless, there are some general patterns that emerge via bioarchaeological
study of past human populations, which this chapter will discuss, in part.
Workload and activity have enormous implications for demographic history of a popula-
tion. The study of living humans indicates, for example, that demanding physical activity in
reproductively aged females results in reduced ovarian function and fecundity (Dufour, 2010;
Ellison, 1994; Jasienska, 2010). Thus, the identification of workload and patterns of physical
activity from the study of human remains may provide indirect reflections of variation in
birthrates and fertility in some past populations.
Of course, we cannot observe and document levels of ovarian function in past populations
directly. However, human remains offer a fund of data for documenting and inferring patterns
and level of workload and other aspects of lifestyle that involve physical activity. Specifically,
the study of pathological and nonpathological changes of articular joints and behaviorally
related modifications of nonarticular regions offers important insights that are not available
from any other record derived from archaeological settings.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 179

5.2 Articular joints and their function


Two types of articular joints, amphiarthrodial (syndesmoses) and diarthrodial (synovial), are
important for interpreting pathological and other modifications in a behavioral context.
Amphiarthrodial joints are somewhat mobile but serve primarily to stabilize specific regions
of the skeleton (e.g., pubic symphysis for the anterior pelvis, intervertebral bodies for the
spine). The ends of bones constituting diarthrodial joints (e.g., knee, interphalangeal, elbow)
articulate with each other within a fibrous capsule, and the articular surfaces are covered with
highly lubricated hyaline cartilage. Depending upon the shape of the articular surfaces, the
anatomy of the capsule, and the ligamentous connections across the joint, freedom of
movement is extensive. Thus, in addition to providing some stability, diarthrodial joints
function primarily in mobility roles, such as extension and flexion of the interphalangeal
joints for grasping, and extension and flexion of the knees for walking and running. Like any
other biological material, the components of these tissues deteriorate, expressing general and
specific pathological changes over the course of a person’s lifetime.

5.3 Articular joint pathology: osteoarthritis


Articular joint deterioration involving bone and associated tissues expresses itself through
multiple conditions, causes, and circumstances. The most common degenerative condition is a
group of joint diseases called osteoarthritis (Felson, 2000; Pritzker, 2003). The clinical,
epidemiological, and research literature represent a confusing array of terms, definitions,
and mixed consensus of etiology. All agree that it is a multifactorial degenerative disorder
involving focal, progressive loss of articular (hyaline) cartilage, often accompanied by
marginal (osteophyte) lipping and articular surface deterioration from direct bone-on-bone
contact (Felson, 2000). Osteophyte formation is the most common expression in archaeo-
logical settings, and likely is an adaptive response to joint instability (van den Berg, 1999).
All authorities also agree that osteoarthritis and its manifestations represent a pattern
of responses to various predisposing factors, including both genetic and environmental/
behavioral causes (Corti & Rigon, 2003; Felson, 2000, 2003; Flores & Hochberg, 2003; Issa &
Sharma, 2006; Manek & Spector, 2003; Sharma, 2001; Valdes & Spector, 2008; Zhang &
Jordan, 2008). There is considerable disagreement on the relative importance of factors, but a
mechanical loading environment due to activity features prominently (Block & Shakoor,
2010; Felson, 2000; Hough, 2001; Jordan, 2000; Jordan et al., 1995; Moskowitz et al.,
2004; Radin, 1982, 1983; Radin et al., 1972, 1991). The mechanical influences having most
agreement are excessive body weight and activity (Abbate et al., 2006; Felson et al., 1988;
Hart and Spector, 1993; Melanson, 2007; Sowers and Karvonen-Gutierrez; 2010; Stürmer
et al., 2000). Clinical and epidemiological studies indicate a greater incidence of osteoarthritis
in obese individuals, especially in the weight-bearing joints and most often in the knee and
hip joints (Abbate et al., 2006; Felson et al., 1988; Jordan et al., 2007; Melanson, 2007;
Sharma, 2001; Sowers & Karvonen-Gutierrez, 2010; Stürmer et al., 2000). The mechanical
stress argument is supported by various findings. For example, industrial laborers show
patterns of articular degeneration in relation to particular physical activities in the workplace.
Strenuous lifting by miners causes articular change in the hips, knees, and vertebrae
(Anderson et al., 1962; Kellgren & Lawrence, 1958; Lawrence, 1977); use of pneumatic tools

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
180 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

by ship builders and others results in similar modifications (Lawrence, 1955, 1961); lifting of
long tongs to move hot metals by foundry workers results in degenerative changes in the
elbow (Hough & Sokoloff, 1989); and repetitive activity involving the hands in cotton mill
workers results in different patterns of osteoarthritis (Hadler, 1977; Hadler et al., 1978).
Other findings for manual laborers, farmers, ballet dancers, various types of athletes, and
those who engage in rigorous exercise generally support these observations (Coggon et al.,
1998; Cooper et al., 1994; Croft, Coggon et al., 1992; Croft, Cooper et al., 1992; Felson
et al., 1988; Forsberg & Nilsson, 1992; Lawrence, 1977; McKeag, 1992; Nakamura et al.,
1993; Stenlund, 1993).
Epidemiological findings are providing important corroboration for conclusions linking
mechanical demand and osteoarthritis. Comparisons reveal markedly higher prevalence of
knee and hip osteoarthritis in North Carolina rural populations than in the United States
(primarily urban) population as a whole (hip: 25.1% vs. 2.7% in the 55–64 age cohort:
Jordan et al., 1995). These differences suggest the greater physical demands of the rural
lifestyle in the modern United States. This same pattern of variation is expressed in at least
one archaeological context. In this regard, the comparison of rural and urban populations
from ancient Corinthian Greece reveals a generally higher prevalence in the former than the
latter, reflecting more strenuous physical labor (Rife, 2012). Rife (2012) suggests that the
greater prevalence of shoulder and vertebral osteoarthritis in the rural individuals reflects
their exposure to work involving agricultural labor, such as field labor, tending livestock,
and other behaviors.
The links between physical activity and osteoarthritis are not straightforward, however.
The hand bones of weavers from the Spitalfields, London skeletal series have no more
osteoarthritis than hand bones from the general sample (Waldron, 1994; and see various
citations in Jurmain, 1999). Manual laborers in this series have no more or less osteoarth-
ritis than the population as a whole. These findings and a survey of inconsistencies found
in the epidemiological literature led Waldron to conclude “that there is no convincing
evidence of a consistent relationship between a particular occupation and a particular form
of osteoarthritis” (1994:94). On the other hand, in some unusual circumstances, there
appears to be a pattern of articular modifications that link with known and highly specific
physical activities (Ciranni & Fornaciari, 2003) or age-related degeneration of a particular
joint or joints of the skeleton in relation to a particular activity regimen (Stirland, 2002;
and see Waldron, 2012). Thus, while articular pathology relating to activity offers an
important insight into behavioral characteristics of human populations in a general sense,
the identification of specific activities or occupations from individual remains is the rare
exception to the general rule.
Like other chronic diseases or disorders, the influence of environmental conditions in the
prenatal environment, infancy, and childhood (low birth weight, poor nutritional status) has
been implicated (Jordan et al., 2005; Melanson, 2007; Peterson et al., 2010). Epidemiologists
and anthropologists observe a great deal of worldwide variation in osteoarthritis in relation to
age (Corti & Rigon, 2003). For example, young adults and older juveniles in some human
populations express a relatively high frequency of osteoarthritis (Chapman, 1972; Chester-
man, 1983; Larsen, Ruff et al., 1995; Rojas-Sepúlveda et al., 2008). In urbanized industrial
societies, osteoarthritis rarely occurs before the age of 40 years (Arden & Nevitt, 2006), a
pattern that is clearly different from the archaeological record showing a much earlier age of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 181

onset (Larsen & Hutchinson, 2010). Regardless of cause, the record shows consensus on one
key factor. That is, owing to disability, the economic, social, and behavioral costs of osteo-
arthritis are substantial in light of the cascade of negative health outcomes from loss of
movement (Corti & Rigon, 2003; Sharma, 2001).
In some respects, epidemiological studies provide an important baseline for interpreting
osteoarthritis in past human populations. For example, the very large samples present a
compelling picture of a huge amount of variation. However, much of the literature is
hampered by the focus on the comparison of folk racial groups (African American vs. White
vs. Asian), having little to do with population biology or modern concepts of human variation
(Allen, 2010). Moreover, bioarchaeological and epidemiological studies are not strictly com-
parable in that the latter are almost always based on clinical contexts, either radiological
examinations or patient interviews, which do not identify subtle degenerative changes seen in
the actual skeletal specimens that bioarchaeologists study. Moreover, clinical evaluations
include factors such as joint capsule spacing. Thus, hard tissue changes observed in the
clinical setting are not strictly comparable to those observed in archaeological or other types
of skeletal collections.
The pathophysiology of osteoarthritis is complex and incompletely understood, especially
regarding the relationship between hyaline cartilage and bone changes. Some have argued
that changes in cartilage – including fibrillation or tearing – precede bony responses; others
contend that minute changes in subchondral bone precede cartilaginous changes (Radin,
1982). For archaeological remains, the exact order of tissue response to mechanical stress is
immaterial because regardless of the order of events, the skeletal changes arising from
osteoarthritis are universal, including proliferative exophytic growths of new bone on joint
margins (“osteophytes” or “lipping”) and/or erosion of bone on joint surfaces (Figure 5.1). In
some joints, the cartilaginous tissue covering the articular surface has failed, resulting in
pitting or rarefaction of the surface. In instances where the cartilage has disintegrated
altogether, the articular surface becomes polished due to direct bone-on-bone contact
(Figure 5.2). Because the surface has a glistening appearance reminiscent of ivory, the
polished area is called eburnation. In the hinge joints of the knee and elbow, deep, parallel
grooves may be present on the eburnated surface (Ortner, 2003; Rogers & Waldron, 1995). The
presence of eburnation indicates that although the articular cartilage is missing, the joint was
still actively used at the time of death (Rogers & Dieppe, 2003).
Osteophytes vary from fine tuft-like, barely perceptible protrusions to large projections of
spiculated bone. Even in the extreme, mobile diarthrodial joints do not usually fuse. In spinal
osteoarthritis, the marginal osteophytes of two adjacent vertebrae may unite, thus forming
a bridge of continuous bone. This change (ankylosis) is also accompanied by reduction in
disk space separating the two vertebral bodies, and hence, marked reduction in mobility of
the spine.
Compression or crush fracture of anterior vertebral bodies – an occasional concomitant
of spinal osteoarthritis – gives them a wedge-shaped appearance (Figure 5.3). Additionally,
herniation of the intervertebral disk results in irregular depressions on intervertebral body
surfaces called Schmorl’s depressions (Ortner, 2003; Schmorl & Junghanns, 1971)
(Figure 5.4).
Biological anthropologists, anatomists, and others have systematically collected data on
osteoarthritis for more than a century. Wells referred to osteoarthritis as “the most useful of all

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
182 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

(a)

(b)

Figure 5.1 Pathognomonic indications of osteoarthritis at various joints: deformation of


the shoulder joint visible on both the right scapula (left) and humerus (right) (a); new
bone growth, marginal osteophytes, and eburnation on the humerus (left), ulnae (center),
and radius (right) (b); marginal osteophytes and pitting at the wrist (c); eburnation and
osteophytosis on the femoral knee surface (d); pitting and osteophytosis of cervical
(e) and lumbar (f) vertebrae; Morropé, Peru. (From Klaus et al., 2009; reproduced with
permission of the authors and John Wiley & Sons, Inc.) (A black and white version of these
figures will appear in some formats. For the color version, please refer to the plate section.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 183

(c)

(d)

Figure 5.1 (cont.)

diseases for reconstructing the life style of early populations” (Wells, 1982:152). Osteoarthritis
is present in all human populations, and regardless of etiology, the patterns documented and
interpreted by bioarchaeologists provide a picture of the cumulative effects of mechanical
stress and age on the body in different human groups. Owing to the lengthy history of study

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
184 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

(e)

(f)

Figure 5.1 (cont.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 185

Figure 5.2 Distal right humerus showing eburnation (osteoarthritis); anatomical specimen.
(From Larsen, 1987; photograph by Barry Stark; reproduced with permission of
Academic Press, Inc.) (A black and white version of this figure will appear in some formats.
For the color version, please refer to the plate section.)

as well as to its ubiquity in skeletal samples, there is a voluminous literature on frequencies


and prevalences in both living and past human groups (Bridges, 1992). There is also
considerable disagreement among paleopathologists about the diagnosis and meaning of
osteoarthritis in archaeological remains. On the one hand, some authorities argue that it is
present only when eburnation is clearly manifested. Alternatively, the presence of at least
two other pathological conditions (i.e., marginal lipping and surface porosity) may be
considered diagnostic (Waldron, 2009). Others, however, regard it as present if any of the
visible pathological changes described in the biomedical/pathology literature are visible
(Hemphill, 2010). The former is influenced by clinical contexts (the visible manifestations
viewed indirectly via various means of imagery), whereas the latter are influenced by
evidence provided from the archaeological context (the visible manifestations via observa-
tion of the bone directly). Moreover, some have downplayed the significance of osteoarth-
ritis in relation to lifestyle reconstruction. Indeed, it is usually not possible to reconstruct
specific habitual activities from the osteoarthritic patterns. However, looking at the general
picture of lifestyle by observation of multiple joints, prevalence, and severity – all within
the context of population and demographics – provides important perspective. The central
point is that this data set can provide meaningful perspective on the relative demands of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
186 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

Figure 5.3 Collapsed thoracic vertebrae; Cochiti, New Mexico. (Photograph by Clark
Spencer Larsen.) (A black and white version of this figure will appear in some formats. For
the color version, please refer to the plate section.)

particular lifestyles. For example, it is highly unlikely that sedentary United States popula-
tions in the twenty-first century exhibit the same profile of osteoarthritis as Great Basin
foragers.

5.3.1 Population-specific patterns of osteoarthritis


Early hominins
Osteoarthritis is present in the earliest hominins, providing an important perspective on
activity patterns in the remote past. The three-million-year-old Hadar australopithecine,
A.L. 288–1 (“Lucy”), displays a distinctive anterioposterior elongation of thoracic vertebral
bodies, marginal lipping, disk space reduction, and intervertebral disk collapse (Cook et al.,
1983). These modifications reflect an extraordinarily demanding activity repertoire, including

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 187

Figure 5.4 Schmorl’s depression on the superior surface of the thoracic body;
Lambayaque, Peru. (Photograph by Sam Scholes.) (A black and white version of this figure
will appear in some formats. For the color version, please refer to the plate section.)

lifting and carrying. The conspicuous anterioposterior vertebral body elongation may be
caused by various activities that involve extreme ventral flexion of the body trunk.
A number of Neanderthal skeletons have distinctive patterns of osteoarthritis that are
useful for reconstructing posture and activity in the late Pleistocene, providing a context
for interpreting behavioral antecedents to modern humans. Based on his study of the La
Chapelle-aux-Saints skeleton, Boule reconstructed the individual “as an almost hunchbacked
creature with head thrust forward, knees habitually bent, and flat, inverted feet, moving along
with a shuffling, uncertain gait” (Straus and Cave, 1957:348). This image of Neanderthal
locomotion served as a model for behavioral reconstruction, and it reinforced the popular
image of Neanderthals as less than human. Straus and Cave (1957) suggested that Boule
misinterpreted key aspects of the anatomy of the skeleton and overlooked the possibility that
severe osteoarthritis may have prevented the individual from normal perambulation. Analysis
of the La Chapelle skeleton reveals the presence of widespread degenerative pathology
(especially marginal lipping) involving the temporomandibular joint, the occipital condyles,
lower cervical vertebrae, and thoracic vertebrae (the T1–T2 exhibits eburnation, and the T6,
T10, and T11 have possible eburnation) (Trinkaus, 1985). The left acetabulum shows extreme
lipping and eburnation. Although the right acetabulum is missing, the head of the right femur
is normal, suggesting that the hip osteoarthritis is unilateral. The severe osteoarthritis in the
left hip suggests that it would have been painful for the individual to walk or run. The overall

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
188 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

pattern of degenerative pathology indicates that locomotor abilities may have been somewhat
limited, but certainly not in the manner imagined by Boule (Trinkaus, 1985).
Osteoarthritis is also extensive in the Neanderthal adults (n¼6) from Shanidar, Iraq (Trin-
kaus, 1983). The widespread nature of articular pathology in these individuals reflects a highly
physically demanding lifeway for these archaic Homo sapiens. This conclusion is confirmed
by other lines of evidence, such as the high overall robusticity and bone strength in these
hominins (Lovejoy & Trinkaus, 1980; Ruff et al., 1993; Trinkaus, 1984; Trinkaus & Ruff, 1999,
2012; and see Chapter 6).

Hunter-gatherers in marginal settings: Sadlermiut Eskimos and Great Basin foragers


The most comprehensive and contextually based bioarchaeological study of osteoarthritis is
the investigation of Sadlermiut Eskimo (Southhampton Island, Northwest Territories) skel-
etons by Merbs (1983). Skeletons in this series display a distinctive patterning of degenerative
articular pathology, which generally matches ethnographically documented activities. Adult
males show bilateral osteoarthritis of the acromioclavicular joint, which is involved mostly in
the elevation of the arm, and hypertrophy of the deltoid tuberosity of the proximal humerus.
A number of potential activities might cause this distinctive pattern of articular pathology and
skeletal morphology, but kayak paddling is the most likely. Extreme loading of the shoulder
and upper arm during kayaking likely contributed to this highly specific pattern of osteoarth-
ritis (Merbs, 1983).
Sadlermiut adult females have high levels of degenerative changes in the temporomandib-
ular joint – twice the prevalence of that in males. This pattern suggests heavy loading of the
mandible, especially in women. As documented ethnographically, adult females habitually
soften animal hides with their dentitions, which may contribute to deterioration of this joint
(Merbs, 1983). Both adult females and males have a high prevalence of postcranial osteoarth-
ritis, which reflects their physically demanding lifestyles. For example, widespread and severe
vertebral osteoarthritis indicates that the backs of both sexes were subjected to marked
compressive forces, such as those that occur during sledding and tobogganing.
Assessment of osteoarthritis prevalence in prehistoric adult males and females from the
American Great Basin similarly contributes to a developing understanding of lifestyle in
marginal settings, especially those that suggest extremely demanding activity and the
ongoing debate about workload in harsh environmental settings (Larsen, Ruff et al., 1995).
In the Great Basin, archaeologists suggest two competing hypotheses regarding subsistence
strategies and resource acquisition generally for the region (Thomas, 1985). One hypothesis
states that prehistoric native populations pursued a limnosedentary exploitive strategy
whereby food and other resources were obtained primarily in ecologically rich, circumscribed
wetland areas that punctuate the desert landscape, thus resulting in a sedentary lifeway.
Alternatively, the limnomobile hypothesis contends that resources in these wetlands do not
provide sufficient resources for the support of native populations, at least on a full-time basis.
These wetlands are subject to occasional resource crashes arising from droughts and floods.
Thus, from this point of view, native populations relied on marsh resources in part, but spent
significant amounts of time in collection and transport of foods recovered from upland
settings in the nearby mountains and elsewhere. The implications of the former hypothesis
is that the more sedentary wetlands-focused adaptation involved less mechanical stress than
the nonsedentary lifestyle; the limnomobile hypothesis is built on the premise that carrying of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 189

supplies and long-distance travel was physically demanding, requiring heightened strength
and endurance of long distance travel (Larsen, Ruff et al., 1995).
In order to determine which of the two models best characterizes adaptive strategies of
native populations in the Great Basin, Larsen, Ruff, and coworkers (1995, 2008) assessed
pattern and prevalence of osteoarthritis in prehistoric human remains recovered from the
Stillwater Marsh region, a large wetlands area in western Nevada. Analysis of these remains
revealed an abundance of osteoarthritis. Most adults, including all individuals over the age of
30 years, have osteoarthritis in at least one, and usually multiple, articular joints. Articular
pathology for older adults involves severe proliferative lipping on joint margins, eburnation,
vertebral compression fractures, and Schmorl’s nodes. Contrary to expectations of the limno-
sedentary model, these findings suggest that hunter-gatherers in this setting led extremely
demanding lives. The high prevalence of osteoarthritis suggests elevated mechanical demand,
such as in heavy lifting and carrying. These findings also imply that prehistoric groups may
not have been tethered to the marsh, but they exploited a wide range of resources from both
the marsh and the surrounding uplands. Beyond concluding that the Great Basin lifeway was
physically demanding, however, it is not possible to state whether these populations were
sedentary or mobile from osteoarthritis evidence alone. Analysis of long bone structural
morphology is more informative on this point (Larsen et al., 2008; Ruff, 2010b; and see
Chapter 6).
Similar assessment of osteoarthritis in the Malheur wetlands in the northern Great Basin
(Oregon) reveals that, like the Stillwater series, there is a strikingly high level of articular joint
pathology, especially in comparison with foragers and later farmers from the Georgia Bight
(Hemphill, 2010) (Figure 5.5). Interestingly, the prevalences for each articular joint, comparing
the two Great Basin series, are statistically indistinguishable (chi-square; P>0.05). The high
values for both series speak to the likelihood that both settings – Stillwater in the western
Great Basin and Malheur in the northern Great Basin – were engaged in very similar adaptive
strategies focusing on largely the same resources, manner of acquiring them, and level
of activity and workload that make the acquisition process economically successful
(Hemphill, 2010).

Population comparisons
These studies of archaic hominins and modern foragers underscore the highly variable nature
of osteoarthritis. In order to assess general patterns of and variation in past physical activities
on a broad basis, comparisons of many different skeletal samples are necessary. Bridges
(1992) attempted a broad-scale analysis by reviewing published studies on appendicular
(shoulder, elbow, hip, knee) and axillary (vertebrae) osteoarthritis in native populations from
North America. In the 25 skeletal samples included in her review, osteoarthritis shows the
highest prevalence in the knee for 17 samples; elbow osteoarthritis is either the most or next
most prevalent for 15 samples. No clear association between osteoarthritis and subsistence
mode emerges in comparing hunter-gatherers and agriculturalists. However, agriculturalists
tend to have a low prevalence in the wrists and hands, but not all foraging groups have high
levels in these joints. For nearly all populations reviewed, ankle or foot arthritis is less
common than hand osteoarthritis.
The comparison of different populations in published findings (Bridges, 1992) contributes
to an understanding of variation in work burdens and activity. However, these comparisons

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
190 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

Males
100 Malheur
Stillwater
Percentage of Individuals Affected
80 Georgia Preagricultural
Georgia Agricultural

60

40

20

–20
Cer Tho Lum Sac Sho Elb Wri Han Hip Kne Ank Ft
Joint Region

Females
90 Malheur

80 Stillwater
Percentage of Individuals Affected

Georgia Preagricultural
70
Georgia Agricultural
60

50

40

30

20

10

–10
Cer Tho Lum Sac Sho Elb Wri Han Hip Kne Ank Ft
Joint Region

Figure 5.5 Frequency and distribution of osteoarthritis among males and females in
the Great Basin (Malheur Lake and Stillwater Marsh) and Georgia coast preagricultural
and agricultural populations.

are limited by the variable nature of the methods of data collection used by the different
researchers (see discussions by Bridges, 1993; Lovell, 1994; Waldron & Rogers, 1991). This
factor alone may prevent investigators from presenting clear diachronic trends or population
differences in osteoarthritis prevalence when comparing findings reported by different
researchers (Cohen, 1989). Data collection and population comparisons by the same researcher

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 191

Table 5.1 Frequency of osteoarthritis in right articular joints expressed by severity


(Adapted from Jurmain, 1980: Table 5)
White Black Pecos Eskimo

Moderate Severe Moderate Severe Moderate Severe Moderate Severe


Joint
Males
Knee 27.0 3.0 38.2 4.5 29.3 1.7 32.4 13.5
Hip 51.0 2.9 47.3 1.8 20.7 2.3 35.2 2.8
Shoulder 47.3 1.1 50.9 3.8 33.3 1.5 53.6 0.0
Elbow 12.5 5.8 19.8 5.2 11.4 3.8 31.1 18.0
Females
Knee 35.6 10.9 31.9 18.6 16.1 0.0 32.0 4.0
Hip 37.4 13.1 47.8 7.8 20.7 0.0 22.4 1.7
Shoulder 44.3 8.2 53.6 8.9 22.2 0.0 23.1 2.6
Elbow 12.7 1.0 21.7 0.9 10.4 3.0 22.0 7.3

or by researchers sharing the same methods circumvent this problem. These types of compari-
sons provide an important perspective on general characteristics of different lifestyles,
especially with regard to workload and level of mechanical demand on the body.
In a classic investigation of variation in degenerative articular pathology, Jurmain (1977a,
1977b, 1978, 1980) assessed osteoarthritis patterns in the appendicular skeleton (shoulder,
elbow, hip, and knee) in a range of populations, including American Whites and Blacks (Terry
Collection), Eskimos (Alaska), and Native Americans (Pecos Pueblo, New Mexico). Eskimos
have a higher prevalence and severity of osteoarthritis than do American Whites and Blacks
or Pecos Pueblo Native Americans; Pecos Pueblo adults have the least prevalence and severity
of osteoarthritis among the four groups (Table 5.1). These population differences reflect the
highly variable mechanical demands associated with contrasting lifestyles and subsistence
strategies. For example, mechanical demands for the Pecos Pueblo agriculturalists may be
mostly limited to the growing season, whereas Eskimos are subjected to high levels of activity
throughout the year (Jurmain, 1977a; and see Merbs, 1983).
The impact of specific lifestyles and occupations on patterns of degenerative articular
pathology in various colonial and postcolonial North American populations has received
considerable attention by biological anthropologists. These studies reveal that for many
settings, physical activities were highly demanding. African Americans from a range of
circumstances provide a growing record of lifestyle and activity in urban and rural settings.
For example, individuals in the First African Baptist Church (urban Philadelphia) cemetery
display extensive degenerative spinal pathology, including osteoarthritis (males, 69%;
females, 39%) and Schmorl’s nodes (males, 31%; females, 13%) (Parrington & Roberts,
1990; and see Angel et al., 1987; Davidson et al., 2002). Similarly, in the African Burial
Ground (urban New York), osteoarthritis is highly prevalent, with the lumbar vertebrae being
especially highly affected (45% of individuals aged 15–24 years; Wilczak et al., 2009). These
prevalences are higher than those of a contemporary African American population from a
rural setting in Cedar Grove, Arkansas (compare with Rose, 1985) and such differences

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
192 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

suggest that the urban lifestyle was far more mechanically demanding than the rural lifestyle.
The differences in degenerative joint pathology between the urban and rural settings may
be due to specific differences in habitual activities. For example, historical records indicate
that individuals interred in the Philadelphia cemetery held unskilled, physically demanding
jobs (see also other settings of African Americans in relation to mechanical environment:
Davidson et al., 2002; Kelley & Angel, 1987; Owsley et al., 1987; Rathbun, 1987; Rathbun &
Steckel, 2002; Thomas et al., 1977).
Highly demanding circumstances are also inferred from the study of osteoarthritis in
pioneer Euroamericans living in the rural frontier of the American Midwest and Great
Plains. Euroamerican adults from Illinois and Texas have remarkably elevated prevalences
of osteoarthritis and highly developed muscle attachment sites on limb bones (Larsen,
Craig et al., 1995; Winchell et al., 1995). Articular degenerative pathology includes
extensive marginal lipping on weight-bearing and nonweight-bearing joints, eburnation,
and extensions of articular surfaces (e.g., anterior femoral head and neck). High prevalence
of nonspecific physiological stress indicators (e.g., enamel defects) and historical evidence
indicate that life on the early American frontier was generally unhealthy and physically
demanding. Numerous historical accounts from the early to mid-nineteenth century
discuss the extremely hard physical labor that pioneer families endured, especially in
preparation of fields and tending and harvesting crops (see discussion in Larsen, Craig
et al., 1995).
Degenerative joint pathology among war casualties is especially revealing about phys-
ical activity in military contexts, mostly drawn from rural, frontier circumstances. Many
of the Euroamerican skeletal remains from the War of 1812 Snake Hill cemetery near Fort
Erie, New York, and the Little Bighorn, South Dakota battlefield display Schmorl’s nodes
on vertebral bodies, with an unusually high prevalence, including the majority of individ-
uals (Owsley, 1991; Scott & Willey, 1997). Some individuals from Snake Hill, for example,
have multiple Schmorl’s nodes: six individuals have five or more vertebrae with nodes,
and one soldier has pronounced nodes in 11 vertebrae. In addition, several individuals
have vertebral compression fractures resulting from excessive mechanical loading of the
back. Similarly, military recruits serving on the mid-sixteenth-century warship, Mary Rose
display a high prevalence of Schmorl’s depressions (thoracic vertebrae: 26.7%) (Stirland,
2002; and compare with Knüsel, 2000). These recruits, composed mostly of adolescents
and young adults, were subjected to heavy mechanical loading in general, and of the
back specifically, in a range of activities, including lifting in quite confined spaces. In
these military settings, the elevated level of vertebral pathology indicate that pre-modern
military recruits were subjected to excessive loading of their spines, such as from lifting
heavy military hardware, carrying heavy loads, and participation in rigorous activity
regimens.
In sharp contrast to these settings, the largely sedentary urban population from
nineteenth-century Belleville, Ontario, presents a remarkably low prevalence of osteoarth-
ritis, including in the spine (Saunders et al., 2002). This speaks to the relatively low amount
of physical activity in comparison with the rural and military individuals described earlier.
In addition, a poorhouse population from Rochester (New York) displays a relatively low
prevalence and severity of osteoarthritis, although not as low as that of Belleville (Higgins
et al., 2002).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 193

Weaponry, food acquisition, and food processing


Ortner’s (1968) classic study of elbow (distal humerus) osteoarthritis in Arctic and Peruvian
Indians (Chicama valley) reveals highly contrasting patterns that reflect different uses of the
upper limb in food acquisition. Arctic populations display a greater prevalence of degenera-
tive changes – marginal proliferation and articular surface destruction – than Peruvian
Indians (18% vs. 5%). Arctic samples also show a distinctive bilateral asymmetry in degenera-
tive pathology; right elbows are far more arthritic than left elbows. Right-sided dominance of
osteoarthritis is due to the greater use of the right arm than the left, such as from spear
throwing with throwing boards (atlatls) by predominantly right-handed hunters (see also
Kricun, 1994; Merbs, 1983; Webb, 1989, 1995).
The prolonged use of weapons over the course of an individual’s lifetime, such as the bow-
and-arrow or atlatl, may also contribute to the degeneration of the elbow joint. Angel (1966b)
first described the “atlatl elbow” in a skeletal series from the Tranquility site, California. He
speculated that the atlatl facilitates a faster spear throw without involving extension and
abduction of the shoulder; extension is primarily limited to the elbow. Consistent with his
hypothesis, Tranquility shoulder joints display very little degenerative pathology, but elbow
osteoarthritis is severe (Angel, 1966b).
In order to document the shift in weapons technology from the atlatl to the bow-and-arrow,
Bridges (1990) assessed patterns of upper limb osteoarthritis in early (Archaic) and late
(Mississippian) prehistoric populations from the Pickwick Basin, Alabama. She suggested that
only one upper limb and specifically the elbow joint is involved in use of the atlatl (Angel,
1966b), whereas both left and right upper limbs and the elbow and shoulder joints of each are
involved in use of the bow-and-arrow. Thus, the joints of the upper limb should show
different distributions of osteoarthritis reflecting either an atlatl pattern (unilateral, elbow)
or bow-and-arrow pattern (bilateral, elbow and shoulder). Because males in most human
societies are responsible for hunting, they should show a higher prevalence of osteoarthritis
than females. As expected, early prehistoric males have a higher prevalence of elbow osteo-
arthritis than late prehistoric males, a pattern that probably reflects the use of the atlatl in the
earlier group and the bow-and-arrow in the later group. Contrary to expectations, both
temporal groups display slight right dominance of osteoarthritis. Early prehistoric females
have the highest frequency of the right-dominant elbow osteoarthritis. These findings provide
mixed support in this setting for the link between weapons use and degenerative articular
pathology.
Angel’s and Bridges’s studies indicate that some groups using the atlatl have a distinctive
pattern of elbow osteoarthritis (e.g., Eskimos), whereas others do not (e.g., Pickwick Basin).
These differences may reflect the relative importance or intensity of specific activities
(Bridges, 1990). For example, traditional Eskimo diets are heavily dominated by meat, and
they relied exclusively (or nearly so) on hunting over the course of the entire year. Thus, their
atlatl use was highly intensive. Early prehistoric Indians living in the Pickwick Basin had a far
more diverse diet that was acquired only partially by hunting. For much of the summer and
spring, native populations utilized riverine resources (e.g., fish) and various flood-plain plants
(e.g., edible seeds); hunting was practiced mostly during the winter. Therefore, the very
different pattern of elbow osteoarthritis in Tennessee Indians cannot be attributed solely to
use of the atlatl or the bow-and-arrow. Rather, a range of activities likely contributed to the
patterns of upper limb osteoarthritis (Bridges, 1990).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
194 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

In contrast to the pattern of right-side dominance of osteoarthritis in upper limbs (Merbs,


1983; Webb, 1995), some groups display bilateral symmetry. Elbow osteoarthritis in native
populations from Chavez Pass, Arizona, is highly prevalent and bilaterally symmetric (Miller,
1985). In this setting, mechanical loading of both elbows while processing maize with
grinding implements – pushing manos against metates with the hands – involves equal use
of the left and right upper limbs (Miller, 1985; and see Merbs, 1980). In traditional South-
western native societies, females are responsible for this activity. Thus, the relatively higher
frequency of such arthritis in adult females in the Chavez Pass series reflects the role of
women in food preparation.

Horseback riding
The horse was an important mode of transport for many Holocene societies, in the Old World
and later in the New World following European contact, and into the early twentieth century.
Some populations show articular degenerative sequelae of an equestrian lifestyle in the
limited number of settings studied by bioarchaeologists (Dutour & Buzhilova, 2014; Edynak,
1976; Larsen, Craig et al., 1995; Owsley et al., 2006; Pálfi, 1992; Reinhard et al., 1994).
Following the introduction of the horse to the American Great Plains by Europeans, native
populations relied on this animal as the key element in the acquisition of resources. Patterns
of osteoarthritis attributed to horseback riding include a high frequency of degenerative
changes in the vertebrae and pelves of adult males in historic-era Omaha and Ponca from
northeastern Nebraska, along with other skeletal features that are best explained by mechan-
ical loading of specific joints during horseback riding (Reinhard et al., 1994). Features
associated with horseback riding are especially diagnostic in the hip joint (innominates,
proximal femora). These features include superior elongation of the acetabulum, extension
of the femoral head articular surface onto the anterior femoral neck (Poirier’s facets), and
hypertrophy of muscle attachment sites for the muscles: adductor magnus, adductor brevis,
vastus lateralis, and gastrocnemius (medial head) (Dutour & Buzhilova, 2014; Erikson et al.,
2000; Reinhard et al., 1994). The development of these hip (and knee) muscles reflects the
emphasis on stabilizing the hip and keeping the rider upright. Extensive osteoarthritis of the
first metatarsals is suggestive of mechanical stresses associated with the placement of the first
toe into a leather thong stirrup (Reinhard et al., 1994). In all settings studied, more males than
females have pathological changes associated with horseback riding, thus indicating that men
were habitually engaged in behaviors involving the use of the horse, more so than women.

Vertebral osteoarthritis
The vertebral column has been studied in a large number of settings in the Americas
(summarized in Bridges, 1992) and elsewhere. For prehistoric North America, these compari-
sons reveal a number of tendencies. First, prevalence is always greatest in the articular region
between the fifth and sixth cervical vertebrae; second, there is a tendency for the lower
thoracic to be affected more than the upper thoracic vertebrae; third, the second to fourth
lumbar vertebrae usually show the greatest degree of marginal lipping in comparison with
other vertebrae; and finally, the region encompassing the seventh cervical vertebra to the
upper thoracic vertebrae (to about the third thoracic vertebra) is always least affected by the
disorder (Bridges, 1992). The relatively minimal amount of osteoarthritis in the thoracic
vertebrae is due to the lower degree of movement in this region of the back (Waldron, 1993).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 195

For a wide range of populations globally, the highest prevalence of vertebral osteoarthritis
is in the lumbar spine, followed by the cervical spine (Bennike, 1985; Bridges, 1994; Gunness-
Hey, 1980; Jurmain, 1990; Klaus et al., 2009; Merbs, 1983; Snow, 1974). Some human
populations show relatively higher levels of osteoarthritis in the cervical vertebrae. For
example, cervical vertebral osteoarthritis is relatively elevated in the Spitalfields, London
industrial urban group (Waldron, 1993). Similarly, Harappan populations from the Indus
Valley display higher frequencies of osteophytes and articular surface pitting of cervical
vertebral bodies than in either the lumbar or thoracic spine (Lovell, 1994). This pattern
suggests an activity-related cause, such as carrying of heavy loads on the head. Individuals
in traditional agricultural communities and from lower socioeconomic groups from urban
settings in South Asia habitually carry loads on their heads (Lovell, 1994). These loads include
laundry bundles, water jars, firewood, and dirt-filled containers at construction sites. Clinical
and observational studies confirm that the upper (cervical) spine is susceptible to injury and
cumulative degenerative changes by persons carrying heavy loads on their heads (Allison,
1984; Levy, 1968; Lovell, 1994). The greater severity of osteoarthritis in the cervical spine in
women than in men suggests that the practice of burden-carrying with the use of the head is
gender specific. For example, severity of cervical osteoarthritis is greater in adult females than
adult males in the Romano-British Bath Gate populations from Cirencester, England (Wells,
1982; and see Lovell, 1994).

5.3.2 Sexual dimorphism in osteoarthritis


Adult males and females show a wide range of variation in osteoarthritis prevalence in the
prehistoric New World and other settings (Bridges, 1992). Some series show a greater preva-
lence in females than males (Rife, 2012) or no difference (Rojas-Sepúlveda et al., 2008), but in
general, males show a consistently greater prevalence of osteoarthritis than females, regard-
less of subsistence strategy or sociopolitical complexity (Hemphill, 2010; Klaus et al., 2009;
Larsen, 1982; Larsen & Hutchinson, 2010; Novak & Šlaus, 2011; Sofaer Derevenski, 2000;
Waldron, 1992; Webb, 1995; Woo & Sciulli, 2013; and many others). Sex comparisons for
prehistoric foragers from coastal Georgia reveal statistically significant differences between
males and females for lumbar (69.2% vs. 32.1%) and shoulder (10.5% vs. 2.4%) joints
(Larsen, 1982). In later prehistoric agriculturalists, more articulations show significant differ-
ences, including the cervical, thoracic, and lumbar vertebrae, elbow, and knee joints. A similar
pattern of increase in sexual dimorphism has been documented in prehistoric northwest
Alabama (Bridges, 1991a). In this setting, differences in osteoarthritis prevalence bet-
ween Archaic-period males and females are not statistically significant, whereas later
Mississippian-period males have more severe osteoarthritis than females (Bridges, 1991a).
These patterns in Georgia and Alabama do not specifically define behaviors associated with
either sex, but they are suggestive of contrasting patterns of physical activity (see later). The
presence of more significant differences between agriculturalist males and females in both
settings suggests the possibility that sex differences in labor demands were greater in later
than in earlier prehistory. Similarly, comparisons of foragers from Indian Knoll, Kentucky,
with maize agriculturalists from Averbuch, Tennessee, indicate different prevalence of osteo-
arthritis between adult males and females (Pierce, 1987). For example, Indian Knoll males
have significantly greater frequencies of shoulder, hip, and knee osteoarthritis than females;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
196 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

Averbuch males have significantly greater frequencies of osteoarthritis for the shoulder and
hip, but not the knee. This pattern is suggestive of change in workload and activity with the
adoption of agriculture.
Unlike males, agriculturalist females from the lower Illinois River valley have a higher
prevalence of vertebral osteoarthritis than forager females from the same region (Pickering,
1984). These differences are especially pronounced in cervical vertebrae, which may be related
to an increase in mechanical demand in this region of the skeleton with the shift to agriculture
(Pickering, 1984).
Fahlström (1981) identified an unusually high prevalence and severity of shoulder osteo-
arthritis in adult males in the Medieval skeletal series from Westerhüs us, Sweden. Historical
analysis of this population suggests that the high frequency in males reflects work and
activity practices that are exclusive to men, including parrying in sword fighting, spear
throwing, timber cutting, and other activities associated with repetitive, heavy loading of
the shoulder joint (Fahlström, 1981).
Some analyses reveal no appreciable differences between males and females. For example,
males and females in the Dickson Mounds, Illinois series show no differences in prevalence of
appendicular osteoarthritis (Goodman, Lallo et al., 1984; Lallo, 1973). The similarity between
sexes infers that mechanical loading of most articular joints in this setting was broadly the
same in adults regardless of sex, in contrast to most other prehistoric Eastern Woodlands
populations (compare with Bridges, 1992). Similarly, documentation of prevalence of
osteoarthritis in two series of African American adults from nineteenth- and twentieth-
century Washington, DC shows no appreciable differences between adult males and females,
suggesting that labor demands were similar for men and women in this setting (Watkins, 2012).
Two clear trends emerge when examining sex differences (Bridges, 1992). First, where there
are statistically significant differences between males and females, males nearly universally
show a higher prevalence of osteoarthritis than females. Second, when looking at specific
regions of the New World, maize agriculturalists tend to display more sexual dimorphism
in degenerative pathology than foragers. This suggests a difference in behavior leading to
degeneration of articular joints in agriculturalists but not in earlier foragers. The change
in pattern of sexual dimorphism suggests that there was a fundamental shift in the division of
labor once agriculture was adopted (Bridges, 1992).

5.3.3 Age variation


The documentation of age-at-onset of osteoarthritis should provide an indication of when
individuals enter the work force. In the late prehistoric Ledders series from the lower Illinois
River valley, elbow and wrist osteoarthritis commences earlier in females than in males, which
may indicate that women were subjected to the mechanical demands of adulthood earlier than
men (Pickering, 1984). Eskimos have the earliest age-at-onset in comparison with South-
western (Pecos Pueblo) agriculturalists and urbanized American Whites and Blacks (Jurmain,
1977a). These differences reflect the relatively greater mechanical demands on the Eskimos in
comparison with other human populations.
Interpretation of intra- and inter-population differences in osteoarthritis prevalence must
consider age structure as it is such an important predisposing factor. For example, females have
a greater prevalence of osteoarthritis than males in all but three of 16 joints in a series of human

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 197

remains from coastal British Columbia (Cybulski, 1992). Adult females are older than the adult
males in the assemblage. Thus, an unusually high prevalence in females relative to males is
likely due to the difference in age composition rather than variation in mechanical environment.

5.3.4 Social rank and work pattern


There is a clear and growing record that tracks inequality in health and quality-of-life among
hierarchical societies, that shows that societies around the world specifically demonstrate
better health in individuals of greater prestige compared to individuals of lower prestige
(Farmer, 2003; Strickland & Shetty, 1998). The study of human remains from archaeological
contexts provides a crucial perspective for examining the origins and evolution of hierarchy
and inequality, especially as it relates to health outcomes (Cohen, 1998), including outcomes
related to labor and workload (Klaus, 2012). Comparison of osteoarthritis prevalence and
severity between social classes in prehistoric stratified societies suggests that higher-status
individuals were exposed to less demanding activities than lower-status individuals.
Archaeological evidence indicates that Middle Woodland populations in the lower Illinois
River valley were hierarchical and organized on the basis of ascribed (hereditary) statuses
(Tainter, 1980). The hierarchy of different social ranks is clearly displayed in the contrasting
levels of energy expenditure in construction of tombs: a great deal of energy and resources
were devoted to the construction of elaborate tombs for high-status individuals. The highest-
rank graves include individuals who were either interred in or processed through large, log-
roofed tombs located at the centers of individual mounds. Little energy was expended on the
construction of tombs for low-status individuals; graves are simple and unadorned. Analysis
of shoulder, elbow, and knee osteoarthritis in skeletons from the Pete Klunk and Gibson
mound groups reveals that the highest-ranking adults over the age of 35 years display less
severe elbow osteoarthritis than lower-ranked individuals, and high-ranking females have
less severe knee osteoarthritis than females from the other ranks (Tainter, 1980).
Similarly, in the Middle Sicán culture (AD 900–1100) in the Lambayeque Valley on the Pacific
coast of Peru, Klaus (2012) tested the hypothesis that the social hierarchy would display marked
differences in workload generally. This hypothesis was based on mortuary analysis, finding
evidence for a small dominant ruling class that was interred with artifacts associated with
remarkable wealth, and access to food that was obtained through the labor of individuals from
lower social classes. The analysis of osteoarthritis in comparing elite and non-elite individuals in
this setting is striking: non-elite are 3.3 times more likely than the elite to have osteoarthritis of
the shoulder, 7.1 times more likely for the elbow, 3.6 times more likely for the thoracic vertebrae,
3.8 times more likely for the lumbar vertebrae, and 4.2 times more likely for the hip. Thus,
overall, there is a strong association between status and workload in the Middle Sicán period of
coastal Peru. These findings strongly suggest that non-elite individuals were involved in activity
regimens that increased the likelihood of excessive joint loading in a range of motions.

5.3.5 Temporal trends and adaptive shifts


The prior discussion underscores the tremendous range of variation in osteoarthritis preva-
lence and pattern, linking the condition to lifestyle, food acquisition, food preparation, age,
social status, and other circumstances. Comparisons of prehistoric foragers and farmers from

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
198 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

different settings (Jurmain, 1977a, 1977b, 1978, 1980) indicate differences in osteoarthritis –
and presumably workload and activity – in relation to subsistence. Regionally based temporal
studies of osteoarthritis give additional perspective on change in functional demand as
populations underwent adaptive shifts in the past. Based on comparisons of earlier and later
societies from the same region, it has become possible to assess the relative labor costs of
change in economic focus, at least as these costs are measured by mechanical stress.
The most extensive temporal studies of osteoarthritis have been completed for several
settings in North America. The study of osteoarthritis prevalence in Archaic-period hunter-
gatherers and later Mississippian-period maize agriculturalists from northwestern Alabama
suggests changes in activity and workload, especially when viewed in the context of diet
and lifeway (Bridges, 1991a). Archaic-period populations exploited a range of terrestrial
and riverine animals and plants, including deer, raccoon, beaver, fish and shellfish, wild
plants, and limited cultivation of sunflower, sumpweed, chenopod, squash, and bottle gourd
(Dye, 1977). Populations moved seasonally from river valleys to nearby uplands. Later
prehistoric groups were intensive maize agriculturalists, but also exploited a limited number
of species of nondomesticated plants and animals (Smith, 1986). These later groups were
largely sedentary and lived primarily in villages on river floodplains, although smaller
temporary upland habitations were utilized for hunting deer and other animals on a
seasonal basis (e.g., small mammals, turkey, waterfowl). In summary, although sharing
some features, the subsistence strategies and settlement patterns in the earlier and later
periods were very different. Because foraging and farming involved very different kinds of
physical activity, the respective populations should display different prevalence and pat-
terns of osteoarthritis.
Comparisons of shoulder, elbow, wrist, hip, knee, and ankle osteoarthritis show a number of
important temporal trends in the Alabama series (Bridges, 1991a). For individuals 30–49 years
of age-at-death, the Archaic group generally has more osteoarthritis than the Mississippian
group, and these differences are especially consistent for males (Table 5.2). Statistically
significant differences between periods are present in only a few of the joints. However, the
overall greater prevalence in the Archaic sample is clear. The severity of osteoarthritis tells the
same story: Archaic populations generally have greater severity of the disorder than Mis-
sissippian populations. The pattern of degenerative pathology is remarkably similar in the
prehistoric foragers and farmers and in the males and females within each group – for all
samples, osteoarthritis is most common in the elbow, shoulder, and knee, and it is least
common in the hip, ankle, and wrist.
Prehistoric and contact-period human remains representing a temporal succession of
Native American populations living in the Georgia Bight have been the focus of research on
physical activity and behavioral changes by Larsen and coworkers (Fresia et al., 1990; Larsen,
1982, 1984, 1998; Larsen & Ruff, 1991, 1994, 2011; Larsen et al., 2007; Ruff & Larsen, 1990;
Ruff et al., 1984). Temporal comparison of osteoarthritis prevalence shows a distinctive
decline in prehistoric farmers relative to earlier foragers. For the series as a whole (sexes
combined), statistically significant reductions occur for the lumbar vertebrae (26.2%), elbow
(6.8%), wrist (4.5%), hip (3.8%), knee (7.2%), and ankle (4.0%) joints. Frequency of osteoarth-
ritis either reduces or does not change in all other joints. Both adult females and adult males
show the same trend of reduction; more significant reductions occur in females than in males
(six joints versus three joints; and see earlier). The pattern of osteoarthritis prevalence in the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 199

Table 5.2 Percentage of individuals with moderate to severe osteoarthritis, aged 30–49 years
(Adapted from Bridges, 1991a: Table 2)
Males Females

Archaic Mississippian Archaic Mississippian

Left Right Left Right Left Right Left Right

Joint
Shoulder 36.8 42.1 30.0b 30.4 7.7 28.6 10.0 17.6
(n) (19) (19) (20) (23) (13) (14) (20) (17)
Elbow 27.3 40.9 28.0b 24.0 26.4 37.6 15.8 20.0
(n) (22) (22) (25) (25) (19) (16) (19) (20)
Wrist 9.5 15.8 0.0 17.4 0.0 6.7 5.6 0.0
(n) (21) (19) (23) (23) (13) (15) (18) (14)
Hip 5.0 5.0 0.0 0.0 0.0 0.0 7.1 0.0
(n) (20) (20) (21) (21) (13) (10) (14) (17)
Knee 27.3 31.8 21.7 8.6 15.8 22.3 21.1 23.5
(n) (22) (22) (23) (23) (19) (18) (19) (17)
Ankle 23.8a–c 0.0 0.0b 4.8b 0.0d 5.9 0.0 0.0
(n) (21) (22) (24) (21) (18) (17) (16) (19)
a
Frequency significantly greater in males than in females (chi-square: P<0.05).
b
Severity significantly greater in males than in females (Mann-Whitney rank sum: P<0.05).
c
Frequency significantly greater in Archaic group than in Mississippian group (chi-square: P<0.05).
d
Severity significantly greater in Archaic group than in Mississippian group (Mann-Whitney rank
sum: P<0.05).

skeleton is similar in both the preagricultural and agricultural groups. In both series the
lumbar and cervical vertebral, elbow, and knee joints show the highest prevalences.
Comparison of the postcontact (seventeenth-century) group with late prehistoric agricul-
turalists reveals a striking increase in osteoarthritis prevalence for most articular joints. Some
of these increases are extraordinary; for example, from 16.3% to 52.9% for the male lumbar
vertebrae and 1.1% to 41.6% for the male foot. Overall, these findings suggest two significant
trends: a decrease in mechanical demand with the introduction of maize agriculture, followed
by a marked increase in mechanical demand following the arrival of Europeans. Foot
osteoarthritis shows the most pronounced increase in comparison with the other articular
joints, especially in males. The increase in degenerative pathology in comparing different joint
types is suggestive of a general and pronounced increase in workload after contact. The very
high increase for the foot suggests that these populations – especially the males – were
engaged in a type or range of activities involving pronounced mechanical demands on the
lower limb. Because the lower limb and foot function primarily in ambulatory activities (i.e.,
walking and running), the increase in contact-era foot osteoarthritis suggests that adult males
were engaged in a great deal of walking.
The changes in osteoarthritis in contact-era populations are consistent with behavioral
characteristics historically documented for the mission period in Spanish Florida. Namely,
native males were drafted into work service under the repartimiento labor system and forced

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
200 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

to make long-distance trips to various localities in the province (Hann, 1988; Worth, 1995).
These trips involved carrying heavy burdens over great distances (Hann, 1988), which placed
demands on the lower limbs in walking, and on upper limbs and trunks. The general increase
in mechanical demands on native populations in the seventeenth century is also well
documented in the historical literature. The Spanish viewed native populations as an inexpen-
sive labor source. Native labor was a central element in their economic and political success in
the area. Indian laborers were used for cargo bearing, agricultural production, construction
projects, woodcutting, and a variety of other physically demanding activities (Hann, 1988;
and see Larsen, 1990a; Worth, 1995). For example, Governor Canzo remarked in his report to
the Crown in 1602–1603 that:

. . .but with all this and the grain from maize, the labor that they endure in the many cultivations that are
given is great, and, if it were not for the help of the Indians that I make them give, and they come from
the province of Guale, Antonico, and from other caciques, it would not be possible to be able to sow
any grain. . . (Unpublished translation provided by J. H. Hann; cited in Larsen, 1990a:16.)

These historical accounts strongly suggest that the increase in osteoarthritis prevalence during
the mission period was due, at least in part, to the increasing labor demands placed on native
groups at that time (see Chapter 6). Similar dramatic increases in oetoarthritis prevalence in
other regions of the New World coming under Spain’s control have been well documented
(Klaus et al., 2009; Ubelaker & Newson, 2002), changes almost certainly due to labor
exploitation (Figure 5.6).
In prehistoric New World populations that underwent the transition from foraging to
farming, there are some trends that point to different lifestyles associated with each adaptive
regime. Both in the Georgia Bight and the upper Midwest (Ohio), there is a decline in prevalence
of osteoarthritis, suggesting a change in lifestyle (Larsen & Ruff, 2011; Larsen, Crosby et al.,
2002; Sciulli & Oberly, 2002). The relatively low prevalence of osteoarthritis in the Georgia
Bight may be influenced by terrain. That is, the comparison of late prehistoric maize farmers
from upland, interior Georgia populations with coastal Georgia populations reveals higher
prevalence of osteoarthritis in the uplands groups (Larsen et al., 2007; Williamson, 2000).
In summary, based on the osteoarthritis evidence, there was an apparent decline in physical
demands with the transition to agriculture in northwestern Alabama and coastal Georgia. It is
important to point out that the Alabama and Georgia populations may not be directly
comparable. There is a large temporal gap between the early (foragers) and late (farmers)
prehistoric groups in northwestern Alabama. The skeletal series used for making comparisons
are separated in time by about 2200 years (1000 BC–AD 1200). The osteoarthritis profile
during the period of time immediately prior to the Mississippian agriculturalists is unknown. It
is possible that a period of less intensive maize agriculture prior to AD 1200 produced levels of
osteoarthritis similar to either the foragers or the farmers. Larsen’s (1982) study of osteoarth-
ritis in the prehistoric Georgia Bight also involved the analysis of skeletal remains from
populations spanning a lengthy period; the precontact preagricultural group is comprised of
all human remains predating AD 1150. However, most prehistoric forager remains are drawn
from the 450-year period immediately preceding the adoption of maize agriculture (c. AD
700–1150). Unlike the Alabama series, there is a cultural and biological continuum with little
or no time gap separating the population groupings. The temporal differences between the
Alabama and Georgia skeletal assemblages may not be significant because subsistence

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 201

(a)
80

Foragers
70
Farmers
60
Osteoarthritis prevalence (%)

Late Mission

50

40

30

20

10

0
C-spine

T-spine

L-spine

Shoulder

Hip

Wrist

Hand

C-spine

T-spine

L-spine

Shoulder

Hip

Wrist

Hand
Males Females

(b)

80 Late pre-Hispanic

70 Postcontact
OA Prevalence (%)

60

50

40

30

20

10

0
C-spine

T-spine

L-spine

Shoulder

Hip

Wrist

Hand

Elbow

Schmorl’s nodes

Knee

Ankle

Foot

C-spine

T-spine

L-spine

Shoulder

Hip

Wrist

Hand

Elbow

Schmorl’s nodes

Knee

Ankle

Foot

Males Females

Figure 5.6 Prevalence and distribution of osteoarthritis in the Georgia Bight (a) and
Lambayaque Peruvian (b) populations, demonstrating the physiological effect of colonial
rule on local inhabitants. Note the peak in osteoarthritis in late mission versus preceding
forager and farming populations, as well as the increase from late pre-Hispanic to
postcontact populations in Peru.

reconstruction based on stable carbon isotopes for the Eastern Woodlands indicates that the
shift to maize agriculture was widespread in the region after about AD 900 (Ambrose, 1987;
Smith, 1989). Thus, the two regions are broadly comparable, at least with respect to the timing
of the introduction of maize agriculture and underlying socioeconomic factors.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
202 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

The adaptive systems represented in the Alabama and Georgia populations are different in
some other important aspects. The Alabama skeletal series represents populations who were
intensive terrestrial maize agriculturalists, whereas the Georgia series represents populations
who were maize agriculturalists, but also engaged in fishing and collecting marine resources
from local estuarine and ocean contexts. Activity differences reflecting these adaptive con-
trasts are suggested by long bone structural analysis (see Chapter 6).
It would be overly simplistic to say that the shift to or intensification of agriculture in
prehistoric North America involved a reduction in workload in native populations. For
example, comparisons of less intensive with more intensive agriculturalists from the
Dickson Mounds site, Illinois show a general increase in prevalence and severity of vertebral
osteoarthritis in the more intensive agriculturalists (Lallo, 1973). In adults (sexes combined),
the frequency increases from 39.7% in the Late Woodland period to 65.8% in the Middle
Mississippian period. Similarly, there is a much greater frequency of osteoarthritis in
Mississippian-period agriculturalists from the Averbuch site compared to foragers from the
Indian Knoll (Pierce, 1987; and see Hodges, 1989).
Much of the focus on the temporal comparisons of osteoarthritis deals with the shift from
foraging to farming. The study of osteoarthritis in relation to other changes in economic
systems has also proven highly informative. In the Santa Barbara Channel islands and
mainland Pacific coast, the focus on hunting and gathering of terrestrial resources was
replaced by intensive fishing in later prehistory (Walker & Hollimon, 1989). The latter
adaptation is especially well documented by early explorers and others who first arrived in
the region. These observations provide important perspective on the types of activities
undertaken by native populations that may have potential influence on articular pathology.
Early accounts of native groups note the presence of an elaborate fishing technology and
material culture, including such items as harpoons, fish traps, nets, and fishhooks. In addition
to fishing, shellfish were collected from rocks by the use of prybars constructed from wood or
bone. Boats made from carved planks were used for travel between islands and between
islands and the mainland. Plant foods (e.g., acorns, chia, and other seeds) were collected in
large quantities, especially on the mainland. Various roots and bulbs were extracted from the
ground with digging-sticks. Economic tasks followed a strict division of labor by gender –
men hunted and fished, and women collected plant foods and shellfish. These differences in
work activities are reflected in dietary differences between adult females and males (Walker &
Erlandson, 1986). For example, early prehistoric women have higher caries rates than early
prehistoric men, which reflects the greater consumption of cariogenic plant carbohydrates by
women (see Chapter 3).
Comparison of osteoarthritis in early and late prehistoric Indians from this setting reveals
temporal changes that are suggestive of alterations in activity and workload with the transi-
tion to a marine focused economy (Walker & Hollimon, 1989). Severity of osteoarthritis,
ranging from slight articular surface porosity to extensive marginal lipping and eburnation,
increases in the late prehistoric period, especially in the lower limb (Walker & Hollimon,
1989). Based on archaeological evidence, Walker and Hollimon (1989) speculate that the
increase in osteoarthritis may be due to increased work that involved trade, exchange, and
more pedestrian travel.
Severity of elbow and wrist osteoarthritis increased in Late period adult males, but not in
adult females in the Santa Barbara Channel Islands region. Osteoarthritis severity declined in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.3 Articular joint pathology: osteoarthritis 203

the shoulder and hand. These changes may be linked to an alteration in weaponry (replace-
ment of the atlatl by the bow-and-arrow) and fishing equipment (shift from harpoons to nets).
The temporal increase in forelimb osteoarthritis is perhaps related to the increased use of
canoes and fishing nets in the Late period (Walker & Hollimon, 1989). Overall, the increase in
osteoarthritis was greater for males than females. Although there are several possible explan-
ations for this trend, the greater role of men in fishing suggests that the workload increase was
greater for males than females. The net result was a decrease in difference in osteoarthritis
prevalence between men and women in the Late period (Hollimon, 1991).
Viewed in a regional perspective, these North American studies show a tremendous range of
variation in mechanical stress loads. This high degree of variability from one setting to
another suggests that at least some factors that influence osteoarthritis are dependent on
local circumstances. These regional studies serve to underscore the point that what often are
perceived as uniform adaptations – even over very large areas – are, in fact, highly variable.
Comparative analysis of osteoarthritis prevalence reveals very different patterns and mech-
anical stress levels. For example, in Australia, elbow osteoarthritis is virtually nonexistent in
east coast populations, whereas it is commonplace in the Murray River valley (Webb, 1989,
1995). This variability reflects a remarkable degree of diversity in use of the upper limb.
The results of analyses of skeletal series representing populations that shifted economic
focus from foraging to farming are available for a limited number of regions in the Old World
(e.g., Europe: Meiklejohn et al., 1984; South Asia: Kennedy, 1984). These studies indicate a
reduction in osteoarthritis prevalence, and suggest a decline in mechanical loading with the
adoption of agriculture, patterns similar to some regions of North America (e.g., coastal
Georgia).
Temporal comparison of earlier and later hunter-gatherers reveals considerable variation,
likely reflecting changes in resource utilization and lifestyle practices. Comparisons of a series
of foragers from the Lake Baikal region of eastern Siberia from the Bronze Age to the Early
Neolithic (c. 6000–1000 BC) reveal key trends, interpreted within the context of a highly
detailed archaeological record of ecology, diet, resource acquisition, and lifestyle (Lieverse,
Weber et al., 2007, 2011; Weber et al., 2010). In particular, analysis of osteoarthritis preva-
lence and patterning provides important insights into activity, workload, and mobility of
foraging populations from a climate having extensive, cold winters and short, mild summers
combined with a wide spectrum of plants and animals. For the 6000-year period, climate and
biota appear to have changed very little. On the other hand, the earlier populations consumed
more fish than the post-hiatus populations (Katzenberg et al., 2012; Lieverse et al., 2011).
Until the final period, there is a slight increase, although not statistically significant, in the
number of individuals affected by osteoarthritis, especially in males. Similarly, the prevalence
and distribution of osteoarthritis in the skeleton is broadly similar for all periods and between
sexes within each period. However, earlier males have a significantly greater prevalence of
knee osteoarthritis than earlier females and later males. These differences may reflect mobility
change over time but with no real change in level of activity, a finding consistent with
archaeological and isotopic reconstructions of mobility in this setting (Lieverse et al., 2011).
One of the most abundant skeletal records for any period and any place is the Medieval
period of the United Kingdom, representing 2635 skeletons dating to three periods, pre-
Medieval (AD 400–1050), Medieval (AD 1050–1550), and post-Medieval (AD 1550–1850)
(Roberts & Cox, 2003). No change in the distribution of osteoarthritis is found in the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
204 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

pre-Medieval and post-Medieval period (Waldron, 1995). However, there is a clear change in
the post-Medieval period relative to the Medieval period. That is, there is an increase in the
proportion of knee osteoarthritis (4.4%) and a decrease in that of hip osteoarthritis (2.9%).
Waldron (1995) suggests that the changes are so rapid that they are probably behavioral in
origin. The pattern of greater knee osteoarthritis continues to the present, suggesting
that the increased incidence of greater body weight – a hugely important predisposing
factor, at least for the knees – may have had its origin centuries ago (see discussion in Roberts
& Cox, 2003).

5.4 Nonpathological articular modifications


Nonpathological skeletal modifications reflecting habitual postures provide a picture of
behaviors such as squatting and kneeling (K. Kennedy, 1989; Trinkaus, 1975). For example,
the crouched posture that characterizes squatting involves extreme flexion of the hip, knee,
ankle, and foot joints. As a result, mechanical demands on lower limb articular joints may
produce distinctive joint modifications. Charles (1893) described an extension of the articular
surface from the femoral head onto the superioanterior neck – called Poirier’s facet – which he
assumed was related to abduction and hyperflexion of the thigh during squatting. However,
the majority of later studies show no relationship between the articular extension and
squatting (Angel, 1964; Trinkaus, 1975). Other skeletal modifications purported to be linked
with squatting include facets on the superioposterior margins of the femoral condyles,
rounding of the posterior aspect of the lateral tibial condyle, retroversion of the tibial plateau,
formation of a groove for the posterior cruciate ligament (on the femoral intercondylar line),
angulated facets on the anterior aspect of the distal tibia, talar neck facet surfaces, and various
other alterations on the talus and calcaneus (reviewed in Trinkaus, 1975). Trinkaus (1975)
concluded that none of these features bears an unambiguous association with squatting. For
example, the presence of articular facets on the superioposterior femoral condyles and the
groove created by the posterior cruciate ligament on the femoral intercondylar line are not
necessarily associated with squatting. The presence of a relatively high frequency of femoral
condylar facets, and extensions in the knee, ankle, and subtalar articulations in western
European Neanderthals is suggestive of a habitual squatting posture. The extreme degree of
skeletal robusticity in these late archaic Homo sapiens indicates very high levels of activity
generally, which would also promote the development of these articular joint modifications.
Therefore, the combination of squatting and great physical activity best explains these
skeletal features, at least in these late Pleistocene hominins (Trinkaus, 1975).
A more diagnostic suite of skeletal adaptations to squatting is tibial retroversion or the
posterior inclination of the tibial plateau (proximal tibia) and presence of lateral squatting
facets on the talus (Boulle, 2001). In a habitual squatting posture where the upper and lower
legs are hyperflexed and the foot is hyperdorsiflexed, a lateral squatting facet develops,
reflecting direct articulation with the anterior rim of the distal tibia. Clear temporal trends
in prevalence of these features show a continuous decline in France from the first to the
eighteenth centuries AD. The decline is especially pronounced to the terminus of the Middle
Ages, suggesting that patterns of resting posture changed dramatically at that time. A rich
historical record links patterns of activity, including the disappearance of an open fire for
cooking and heating and the wider availability of furniture for sitting (especially chairs and
tables), to the alterations in resting postures occurring at that time.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.4 Nonpathological articular modifications 205

Trinkaus’s (1975) study reminds us that it is difficult to sort out general activity levels and
specific articular joint modifications in interpreting skeletal morphology. The study of dis-
tinctive metatarsophalangeal joint articular modifications in several different contexts reveals
that habitual kneeling – in food preparation tasks or other occupational activities – can be
identified. When walking or running, a high degree of dorsiflexion at the metatarsophalan-
geal joint of the toes is sustained only momentarily. In kneeling postures where the hyper-
dorsiflexed position of the toes is sustained for extended periods, the joints develop articular
modifications reflecting these behaviors. Metatarsophalangeal alterations have been identi-
fied in late prehistoric (AD 700–1550) human remains from the Ayalan site, Ecuador (Ubela-
ker, 1979). The articulations are characterized as small extensions or facets, or both, on the
distal ends of metatarsals and the proximal ends of proximal foot phalanges (Figure 5.7). The
facets are flat with clearly demarcated proximal borders and an accompanying superior
surface bony extension extending distally from the proximal articular surface. These alter-
ations are present in about 20% of Ayalan foot bones; they are bilaterally distributed among
the first three metatarsals and first phalanx, and are far less common in the fourth and fifth
metatarsals.
Comparisons of the Ayalan series with other archaeological series (Eskimos, Hawikuh site
Zuni, Mobridge site Arikara, Late Woodland Nanjemoy, and Terry Collection Blacks and
Whites) show a great deal of variation in metatarsophalangeal articular facets (Ubelaker,
1979). The alterations occur in all of these samples, but they are most common in the Ayalan
group. The prevalence in the Hawikuh and Nanjemoy samples is 5%; the other groups have
prevalences of 2% or less. The distribution among the digits of the foot also varies between
samples. In the Ayalan series, the alterations are present on the first through fourth metatar-
sals and first proximal phalanges, whereas in the Hawikuh and Eskimo series, they are on the
second through fourth metatarsals.
Although little research has been conducted on these articular modifications, they are
found in a wide diversity of human populations. In addition to those analyzed by Ubelaker
(1979), Mesolithic and Neolithic skeletal series from the early agricultural settlement at Tell
Abu Hureyra, Syria, display metatarsophalangeal alterations in the first metatarsals (Molle-
son, 1989, 1994). In older adults, the margins of these facets are associated with degenerative
changes (osteoarthritis). At Bronze-Age Bab edh-Dhra’12, Jordan, adults display more alter-
ations in the earlier than later occupation, suggesting alteration in activity involving a change
in repetitive activity and posture (Ullinger et al., 2012). Proximal extensions of the distal
articular surfaces in many of the metatarsals are also identified in prehistoric samples from
the north coast of Rota, Mariana Islands (Hanson, 1988). The association with significant
mechanical stresses of the foot is also indicated by the co-occurrence of osteoarthritis in this
setting. Similar joint modifications of the first metatarsals are present in nineteenth-century
fur-trappers from Alberta, who in life spent long hours canoeing from one location to the
next, primarily in kneeling postures with their toes dorsiflexed (Lai & Lovell, 1992; Lovell &
Lai, 1994). Likewise, the characteristic morphology, location, and association with osteoarth-
ritis in older adults in Syria and Ecuador indicate that the metatarsophalangeal joint alter-
ations were probably produced by prolonged hyperdorsiflexion of the toes while kneeling
(Molleson, 1989, 1994; Ubelaker, 1979). With regard to the Ecuador series especially, Ubelaker
(1979) speculates that kneeling while grinding maize with stone metates was the most likely
cause. Molleson (1989) provides confirmation of a similar posture in regard to cereal grinding
depicted in Assyrian and Egyptian dynastic tomb art.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
206 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

Figure 5.7 Nonpathological alternations of distal metatarsals produced by


hyperdorsiflexion of the toes; Ayalan, Ecuador. (Adapted from Ubelaker, 1979;
reproduced with permission of author and John Wiley & Sons, Inc.)

5.5 Nonarticular pathological conditions relating to activity


5.5.1 Entheseal changes
An enthesis is a site where a tendon or ligament interfaces with bone. This interface anchors
the muscle and dissipates stress. From a biomechanical perspective, entheses represent critical
concentrations of stress involving the juncture between hard and soft tissue (Benjamin et al.,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.5 Nonarticular pathological conditions relating to activity 207

(a)

Figure 5.8 Entheseal changes on the medial epicondyle of the humerus (a), radial
tuberosity of the radius (b), and attachments for infraspinatous and teres major muscles.
(Photographs by Sébastien Villotte.) (A black and white version of these figures will
appear in some formats. For the color version, please refer to the plate section.)

2006). Various skeletal elements are affected, but enthesial changes are especially common-
place for the humerus, radius, ulna, tibia, femur, metacarpals, metatarsals, and terminal
phalanges (Bufkin, 1971; Edgerton et al., 1990; Owsley et al., 1991).
Although referred to in the literature as musculoskeletal stress markers, cortical defects, or
other synonyms (Brower, 1977; Bufkin, 1971; Edgerton et al., 1990; Owsley et al., 1991;
Resnick & Greenway, 1982), the irregularities on the outer margin of the joint and/or other
characteristics (e.g., foramina) are called enthesial changes (Villotte & Knüsel, 2013)
(Figure 5.8). Although the link between form and function is debated, it is clear that enthesial
changes provide a potentially important source for inferring patterns of activity (Dutour,
1986; Hawkey, 1998; Hawkey & Merbs, 1995; Ibanez-Gimeno et al., 2013; B. Kennedy, 1989,
1998; Niinimäki, 2012; Ruff et al., 2006; Schlecht, 2012; Stirland, 2002; Villotte et al., 2010a,
2010b; Weiss, 2007; various in Henderson & Cardoso, 2013).
This body of work was motivated by a set of investigations undertaken by Angel and his
collaborators, beginning in the 1950s, who were interested in reconstruction of physical
activity from analysis of skeletal morphology (Angel, 1952; Angel et al., 1987; Kelley &
Angel, 1987). There is now an extensive record of research focusing on enthesial changes,
much of which is built on the influential study of Hudson Bay Inuit skeletal remains where
robusticity markers (degree of muscle markings), stress lesions (pitting or furrows), and
ossification exostoses (bony spurs) for each muscle attachment site were documented by four
well-defined grades (Hawkey & Merbs, 1995). The results of this and other investigations
suggested patterns of variation that reflected muscle use and activity involving specific joints.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
208 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

(b)

Figure 5.8 (cont.)

Most investigations following Hawkey and Merbs’s (1995) study use this method or its
derivations (Eshed et al., 2004; Hawkey, 1998; Lieverse et al., 2009, 2011; Molnar, 2006, 2008;
Molnar et al., 2011; Perry, 2008; Peterson, 1998, 2002; Steen & Lanes, 1998; Weiss, 2007).
Approaches to data collection based on understanding of the types of entheses – fibrous and
fibrocartilaginous – provide a more biologically meaningful assessment of enthesial changes
in relation to function and activity (Schlecht, 2012; Villotte & Knüsel, 2013; Villotte et al.,
2010a, 2010b). These two types of entheses have biomechanical and histological differences, a
factor not recognized in most earlier studies using form to infer function. Villotte and
coworkers (2010a, 2010b) propose focusing on fibrocartilaginous entheses, both in order to
understand the underlying etiology to better inform behavioral inference and to avoid
conflating different tissue types and drawing generalizations about activity.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.5 Nonarticular pathological conditions relating to activity 209

(c)

Figure 5.8 (cont.)

Regional variation and interpretation


Proximal right ulnae of adult males from terminal Pleistocene sites in central India show a
high frequency of enthesial changes characterized as well-developed supinator crests (Ken-
nedy, 1983). The insertion areas for the anconeus muscle on the ulna are also unusually
pronounced. These features occur in relatively high frequencies in Mesolithic populations
from Gujarat, Rajasthan, the Ganga plain of Uttar Pradesh, and Sri Lanka (Kennedy, 1983;
Lukacs & Pal, 2003). Kennedy (1983) contends that well-developed supinator crests reflect the
heavy use of missile weapons (e.g., spears, bolas, slings, boomerangs) by South Asian
foragers. This type of throwing involves various movements that involve the supinator muscle
directly, including abrupt shifts from supination to pronation of the forearm. Muscular
strength is the critical factor in throwing abilities and ultimately in the formation of hyper-
trophic supinator crests (Kennedy, 1983; Lukacs & Pal, 2003).
In some Arctic populations, the high prevalence of enthesial changes generally indicates
high levels of physical activity. Human remains from the eastern Aleutian Islands (Akun and
Akutan Islands) have such a high prevalence (Hawkey & Street, 1992). Aleut females show
evidence of marked stress for the right hand and wrist, whereas use of the left arm appears to
have been associated with habitual adduction and abduction. In males, the upper limbs show
bilateral skeletal changes, involving both humeri, resulting from extremely heavy use of left
and right arms. These skeletal modifications may result from endurance kayaking with
double-bladed paddles. Thule Eskimo groups from northwest Hudson Bay have similar
patterns of bilateral upper limb skeletal modifications, indicating the widespread use of
kayaks and the general demands of living in the harsh Arctic setting (Hawkey & Merbs,
1995). Temporal comparisons suggest shifts in adaptive strategies involving seal hunting. For

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
210 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

example, an increase in markers for pectoralis major in comparing Early and Late Thule males
suggests an increase in downward thrusting, such as that associated with harpooning
seals. Ethnographically, this thrusting motion targets the seal as it comes to its breathing
hole in the ice.
Musculoskeletal functional analysis of entheses in other settings where archaeological,
ethnographic, historical, and other contextual evidence is available reveals important behav-
ioral and lifestyle interpretations. On the island of Gotland (Sweden; see Chapter 4), where
Neolithic (3400–2300 BC) coastal populations focused their diets on fish and seals from the
Baltic Sea, markers are higher in males than females and in older adults compared to younger
adults (Molnar, 2006, 2008). The common trend for many of the investigations of enthesial
changes suggests that much of the variation can be explained by accumulation of modifica-
tions with advancing age. This makes the strong case that age must be accounted for when
documenting enthesial changes for reconstructing and interpreting activity patterns (Weiss,
2007).
In contrast to the clear age progression of enthesial changes in the series from Gotland,
Sweden, there are clear age changes, but for females only, in the later post-hiatus series from
Lake Baikal, Russia (Lieverse et al., 2009, 2011). Moreover, there is an increase in the severity
of upper limb changes in the post-hiatus males. These patterns of general increase, especially
in males, are consistent with the notion that later-period populations were involved in
hunting forays, in particular hunting large terrestrial mammals (e.g., ruminants). Interest-
ingly, in both males and females, markers for conoid and costoclavicular ligaments and
pectoralis major, deltoid, pronator quadratus, and brachioradialis were quite pronounced,
regardless of the period.
There is considerable debate about the foraging-to-farming transition in the Levant,
especially in the region encompassing the modern country of Israel (Hershkovitz & Galili,
1990). In order to address the question, did the transition to farming involve reduced
workload? Eshed and collaborators analyzed upper limb markers for four pre-farming (Natu-
fian) and four early farming (Neolithic) skeletal series. Based on the consistent increase in
scores, the results of their study suggest a clear increase in mechanical stress and loading of
the upper limbs. Males generally show greater scores than females in both periods. However,
Natufian females express higher scores than males for the deltoid attachment and show a clear
dominance of right over left side. Eshed and coworkers (2004) suggest that this reflects the
elevated use of the forearm muscles, perhaps in pounding of grain with a stone or wood pestle,
but likely other activities requiring heavy use of the arm in a range of work activities. In her
analysis of changes in the foraging-to-farming transition in the Levant, Peterson (2002)
found a distinctive Natufian pattern of high scores in adult male upper limbs, especially for
the right triceps brachii and anconeus, muscles involved in forearm extension. This kind of
movement is associated with activities in overhand throwing motions, such as for atlatl use
and spear throwing. Some of the individuals were young adult males, suggesting that the
pattern is not just an age-related phenomenon. On the other hand, both males and females
show similarities in scores, such as for pectoralis. This suggests overlapping patterns of upper
limb use for males and females. In sharp contrast, the Neolithic adults show a very different
pattern than the Natufian adults. In particular, Neolithic adults show a distinctive symmetry
in scores comparing right and left sides for both males and females, suggesting bilateral tasks
in the use of the upper limbs. In addition, the scores for the brachialis – a forearm flexor – are

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.5 Nonarticular pathological conditions relating to activity 211

much more pronounced in the Neolithic males and Natufian males. Overall, the pattern of
changes shows fewer sex differences and more emphasis on bilateral activities with the shift
to and intensification of farming. The archaeological context shows emphasis on farming
activities, including field preparation, harvesting, and food preparation.
In contrast to the symmetry of enthesial changes in the upper limbs in early farmers in the
Levant, prehistoric males from Europe express greater frequency of enthesial changes in the
right elbow (humeral epicondyles) than in the left elbow (Villotte & Knüsel, 2014). This strong
degree of asymmetry is consistent with an extensive record of articular injuries resulting from
overhead throwing using the right arm, a largely male-based activity pertaining to use of the
atlatl and other projectile technology.
Analysis of a large agricultural-based population from Grasshopper Pueblo, Arizona
(Pueblo IV period: AD 1275–1600) expresses a very different kind of pattern in comparison
with the agriculturalists from the Levant and some other contexts (Perry, 2008). In this setting,
males show strong unilateral emphasis on muscle groups that are involved in movement of
the arm and shoulder, stabilization of the chest and shoulder, and flexion of the arm. The
right–left asymmetry with right dominance is especially pronounced for the right deltoid and
costoclavicular ligaments in males, and at considerably higher levels than for females. There
are various activities that could explain this highly distinctive pattern in the late prehistoric
Southwest. However, in the context of behavioral reconstruction based on archaeological,
ethnohistorical, and ethnographic evidence, the most likely male activity includes use of the
bow-and-arrow in large game hunting and in its masculine identity. The distinctively bilateral
development of changes in the upper limb in females likely reflects corn grinding and other
domestic activities associated with food preparation. On the other hand, females show a
distinctive unilateral development of the transverse head of the adductor pollicus, a muscle
that brings the thumb to the center of the palm of the hand, such as that associated with
scraping ceramics, using a scraping tool held in the right hand during production. Similarly,
analysis of War of 1812 battle casualties from Snake Hill, Ontario reveals an unusually high
frequency (40%) of enthesial changes on proximal humeri, significant higher than in other
series, such as Civil War casualities and Native American hunter-gatherers and agricultural-
ists (Owsley et al., 1991). Moreover, there is significant right dominance (62.5%) in the Snake
Hill series; if bilaterally present, the defect on the left side is small and shallow compared to
the much larger defect on the right side. This pattern of variation likely reflects right hand
dominance.
Overall, therefore, sex and age patterns of enthesial changes in archaeological settings are
highly consistent: males generally express higher prevalence, severity, and right-dominated
upper limb articular injury than females, which can broadly be interpreted to mean that males
have greater and more intensive physical activity and muscle use than females. Similarly,
with few exceptions, severity increases with age, and the most severe forms of enthesial
changes are associated with elderly adults (Villotte et al., 2010a, 2010b).
One would expect to find evidence for a relatively greater workload and demanding
lifestyle in low-status individuals who had been exploited as a labor source. Indeed, this
expectation is met with regard to exploitation of native populations by Spaniards during the
colonial era in the American Southwest (Chapman, 1997) and in a Medieval-era setting in
east-central Europe (Havelková et al., 2011). In both settings, individuals exposed to a high-
level mechanical demand express elevated prevalence of enthesial changes.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
212 Activity patterns: 1. Articular degenerative conditions and musculoskeletal modifications

Although analyses of enthesial variation from bioarchaeological settings are highly sug-
gestive, there remains very little direct evidence for a relationship between entheses variation
and activity, either specifically involving a particular muscle group or general level of degree.
Most of the aforementioned studies assume that activity is responsible for increasing perios-
teal apposition in those areas of the skeleton associated with muscle tendon insertions
(Schlecht, 2012). That is, bioarchaeologists have extrapolated from the findings of experi-
mental skeletal biologists (Chamay and Tschantz, 1972; Woo et al., 1981) who have docu-
mented the direct link between activity increase and periosteal apposition in tendinous
regions of the skeleton. Applying these findings, bioarchaeologists argue that the periosteum
associated with muscle sees increases in periosteal capillary volume with an increase in
muscle contraction. This causes the periosteum in the region associated with the muscle to
hypertrophy, thereby increasing the strength of the tendon-anchoring sites (Hawkey & Merbs,
1995; Torg et al., 1972; Zumwalt, 2006). Schlecht (2012) notes, however, that while the work
of experimental skeletal biologists has documented some key insights, they may not be
applicable to understanding the etiology and derivation of entheses. Importantly, little work
has been done on enthesial changes, documenting variation in loading in the regions
involved. Analysis of enthesis morphology of exercised sheep compared with non-exercised
sheep in an experimental context revealed no linkage between enthesis morphology and
exercise (Zumwalt, 2006). On the other hand, histomorphological analysis of the proximal
radius in a human cadaver sample of known age and sex revealed evidence of increased bone
turnover, suggesting a potential relationship between enthesial changes and mechanical
strain (Schlecht, 2012). This record – bioarchaeological, experimental, and anatomical –
indicates that enthesis morphology, while likely linked to activity, is complex and incom-
pletely understood.

5.6 Summary and conclusions


The study of articular joint modifications relating to mechanical demand – especially osteo-
arthritis – offers insight into the stresses of different activity patterns and lifestyles in past
populations. Generally, the more mechanically demanding the lifeway, the greater the preva-
lence of osteoarthritis and other degenerative pathological conditions related to bone fatigue
and fragility (e.g., spondylolysis; see Chapter 4). Conversely, less demanding work repertoires
result in a relatively lower prevalence of these conditions. Temporal comparisons of contrast-
ing subsistence strategies within regions indicate some suggestive trends. For example,
although there are some exceptions, later prehistoric agriculturalists tend to show a reduction
in osteoarthritis and degenerative pathology than earlier foragers. This provides some support
for the traditional point of view that foragers work harder than agriculturalists. More
importantly, local circumstances and conditions influence osteoarthritis prevalence and
pattern in a far more profound way. Adult males generally show a tendency for greater
degenerative pathology than adult females in archaeological settings. Although osteoarthritis
is related to mechanical loading, the relationship to level or type of activity is not a direct one.
High levels of osteoarthritis in the skeleton suggest demanding lifestyles, but do not indicate
whether these demands also include long-distance or frequent travel as a cause, except
possibly with regard to a high prevalence of foot osteoarthritis.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
5.6 Summary and conclusions 213

The study of enthesial changes indicates that they are likely related in some manner to
activity and lifestyle, providing an important avenue of inquiry for interpreting behavior as it
relates to activity and lifestyle. Additional exploration, especially through histological analy-
sis, experimental investigation, and covariance with other indicators of activity (e.g., cross-
sectional geometry), have the potential to provide a much clearer understanding of the
adaptive responses to a diversity of activity regimens and variation by sex, age, body size,
and labor intensity (Ibanez-Gimeno et al., 2013; Jurmain, 1999; Niinimäki, 2009; Pearson &
Buikstra, 2006; Schlecht, 2012; Stirland, 1998; Villotte et al., 2010a, 2010b; Weiss, 2007;
Zumwalt, 2006, Zumwalt et al., 2000).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:28:45, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.007
6 Activity patterns: 2. Structural adaptation

6.1 Bone form, function, and behavioral inference


Julius Wolff, a leading nineteenth-century German anatomist and orthopedic surgeon, recog-
nized the remarkable sensitivity of skeletal tissues to mechanical stimuli, especially with
regard to their ability to adjust size and shape in response to external forces. Wolff concluded
that “every particle of mature bone is very active. Such activity must appear in the external
shape of the bones” (1892:78). What he called the “law of bone remodeling” – now commonly
known as Wolff’s Law – simply states that bone tissue places itself in the direction of
functional demand. Wolff’s Law in the twenty-first century has a somewhat different mean-
ing than what Julius Wolff intended, and is best referred to as “bone functional adaptation”
(Ruff et al., 2006).
A great deal of evidence has accrued in support of the notion that bone adapts to and is shaped
by its mechanical environment. Experimental and other research on bone modeling and remod-
eling is instrumental in identifying patterns of skeletal modification under different loading
regimes, especially with respect to the magnitude and direction of mechanical forces (Bass et al.,
2002; Lanyon et al., 1982; Meade, 1989; Ruff, 2008; Trinkaus et al., 1994). In a set of classic
experiments using laboratory dogs, Chamay and Tschantz (1972) observed that the surgical
removal of portions of radii resulted in the hypertrophy of ulnar diaphyses. The ulnar diaphyses
increased in size by 31% after just 16 days and 60%–100% by nine weeks. Similarly, Lanyon and
coworkers (Goodship et al., 1979; Lanyon & Bourne, 1979) documented increased apposition of
bone on the radius following ulnar osteotomies in pigs and sheep. Nonsurgical load alterations
have also resulted in changes in bone mass. Woo and coworkers (1981) identified significant
endosteal apposition in young pigs subjected to exercise. Simkin and coworkers (1989) com-
pared humeri from swimming and non-swimming rats in an experimental setting. The swim-
ming rats included a group trained to swim for one hour per day and a group that underwent the
same training, but also had a lead weight (approximately 1% of the rat’s body weight) tied to
their tails. Comparison of bone size and structure revealed that both groups of swimming rats
had greater periosteal apposition than the sedentary, non-swimming rats.
Humans with unusually high levels of activity involving use of an extra-dominant upper
limb, such as professional racket sports players (Jones et al., 1977; Kannus et al., 1995; Ruff
et al., 1994; Shaw & Stock, 2009a), rodeo cowboys (Claussen, 1982), and baseball pitchers
(King et al., 1969), exhibit marked hypertrophy of the external diaphyses of long bones of the
playing side. In professional tennis players, for example, one study revealed that males have a
35% increase in cortical area in the distal humerus of the playing arm versus the non-playing
arm; females have a 29% increase (Jones et al., 1977). The effects of elevated mechanical
demands in relation to increased exercise are well documented in clinical settings. Adults who

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 215

exercise tend to have higher bone mass and size than adults who are relatively sedentary,
especially in males (Lorentzon et al., 2005).
Loss of normal functional loading, especially involving extended periods of bed rest,
weightlessness in spaceflight and absence of gravitational loading, or from partial or complete
immobilization of limbs, results in decreased bone mass (Abram et al., 1988; Holick, 1998;
Jenkins & Cochran, 1969; Kazarian & Von Gierke, 1969; Kiratli, 1992; Lanyon & Rubin, 1984;
Lazenby & Pfeiffer, 1993; Meade, 1989; Miyamoto et al., 1998; Morey & Baylink, 1978; Peck &
Stout, 2009; Prince et al., 1988; Sevastikoglou et al., 1969; Todd & Barber, 1934; Whalen
et al., 1988; see discussion in Trinkaus et al., 1994). The loss in mass is largely due to profound
reduction in remodeling rates of bone whereby the normal linkages between osteoclasts and
osteoblasts are decoupled (Alexandre & Vico, 2011; Schlecht et al., 2012). This dysfunctional
replacement process leads to more removal, especially endosteally, than replacement of bone,
and ultimately, considerable reduction in bone density and mass.
Bone is anisotropic in that it is characterized by different material properties depending
upon the direction of loading (Cowin, 1989; Currey, 2002; Nordin & Frankel, 2001;
Wainwright, 1988). Long bones, for example, are stronger in the longitudinal direction than
in any other plane (Wainwright, 1988). The primary loading forces affecting bone include
tension, compression, shearing, bending, and torsion (Figure 6.1). These can be best understood
by considering a thin section, or slice of bone. Tensile loading occurs when equal and opposite
forces are applied outwardly from the surface of the slice. Compression, the opposite of tension,
occurs when equal and opposite loads are directed toward the two surfaces. Shearing loads
involve the application of forces parallel to the surfaces under consideration. Bending forces
produce two types of stresses, tension on the convex side and compression on the concave side.
Torsional loading is the twisting of the skeletal element about an axis and results in a
combination of tension, compression, and shear forces. Individual skeletal elements have an
irregular geometric structure, and a range of forces usually acts on bone in normal physio-
logical activities, such as running and walking. Therefore, loading almost always involves a
combination of these modes, but the largest and most common loading parameters affecting
the human skeleton are bending and torsion, especially for the long bones.
This extensive body of research makes clear that there is strong support for inferring patterns
of activity from skeletal morphology. However, it must be kept in mind that there are other
factors that influence morphology. There are genetic influences, and skeletal form represents
compromises between mechanical and other influences, such as nutrition, hormones, and age
(Gosman et al., 2011; Nelson et al., 2004; Ruff, 2008; and see Chapter 2). This chapter takes the
position that mechanical loading, as it relates to lifestyle and habitual activity, plays the leading
role in explaining robusticity and its variation (Daly et al., 2004; Martin, 2007; Robling et al.,
2002; Stock, 2006). In the larger picture, this record reveals the fundamental importance of
adaptive plasticity, or the ability to respond to environmental circumstances during the years of
growth and development (Gotthard & Nylin, 1995; Rovosa et al., 2008; and see Chapter 7).

6.2 Cross-sectional geometry


Biomechanics (sometimes called “mechanobiology”) – the application of engineering principles
to biological materials – represents an important means of analyzing and interpreting skeletal
morphology within the context of the mechanical environment (Carter & Beaupré, 2001;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
216 Activity patterns: 2. Structural adaptation

UNLOADED TENSION COMPRESSION BENDING

SHEAR TORSION COMBINED LOADING


TORSION-COMPRESSION
Figure 6.1 Loading modes that affect long bones. (From Nordin & Frankel, 2001; reproduced with
permission of authors and Williams & Wilkins.)

Currey, 2002; Ruff, 2008). Unlike straight mechanical analysis of building materials, biomech-
anics deals with dynamic tissues that modify themselves continuously in relation to loading
modes and activity levels. That is, building materials do not adapt and alter their form in
response to the mechanical environment. Only dynamic, living tissues adapt and alter their
form in response to the mechanical environment.
Density of bone tissue differs within the skeleton and within individual bones in response to
varying mechanical demands. As such, mineral content is a component of bone strength
(Burr, 1980; Martin et al., 1998). However, the response to increased loading is primarily in the
distribution of bone rather than in its density or any other intrinsic material property (Beck
et al., 1990; various in Burr & Allen, 2013; Burr et al., 1989; Gosman & Ketcham, 2009;
Grynpas, 2003; Ruff, 2008).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 217

Borrowing the simple beam model used by civil and mechanical engineers to analyze
buildings and design structures (Huiskes, 1982; Timoshenko & Gere, 1972; Turner & Burr,
1993), biological anthropologists and others have analyzed stresses – the pressures on the
inside of materials – that result from loads applied externally. These stresses have been
calculated via analysis of cross-sectional geometric properties in various settings, including
archaeological (Bridges, 1989b; Bridges, Blitz et al., 2000; Larsen & Ruff, 1994; Lieverse et al.,
2011; Maggiano et al., 2008; Ruff & Hayes, 1983a, 1983b; Ruff & Larsen, 2001; Stock &
Pfeiffer, 2001; Stock et al., 2011), fossil hominins (Churchill & Smith, 2000; Holt, 2003; Holt &
Formicola, 2008; Trinkaus & Ruff, 1999; Trinkaus et al., 1994), and non-human primates (Burr
et al., 1989; Kimura, 2002; Polk et al., 2000; Ruff & Runestad, 1992). These investigations, as
well as experimental evidence based on laboratory animals (Abram et al., 1988; Simkin et al.,
1989), indicate the value of structural analysis in drawing inferences about physical activity
and behavior patterns. However, experimental research also suggests that mechanical loading
of long bones is an approximation of the loading of simple beams (Lieberman, Polk et al.,
2004), but the large and growing body of work on humans and animals indicates that cross-
sectional geometric analysis is the best available source for understanding variation in
mechanical loading and integrity (Ruff, 2008; Ruff et al., 2006; Turner & Burr, 1993).
Stresses are expressed as two specific characteristics: rigidity (the ability to resist bone
deformation during loading) and strength (the ability to resist structural failure or fracturing
during loading). Both characteristics are central to understanding skeletal adaptation. In
bending and torsion of a hollow beam, such as a long bone, the magnitude of stresses is
proportional to the distance from the central or “neutral” axis of the bone (Figure 6.2). The
neutral axis is the plane (bending) or axis (torsion) where stress is zero. Thus, all else being
equal, the cross-section that is strongest is that in which the material is oriented furthest from
the neutral axis (Currey, 2002; Nordin and Frankel, 2001; Ruff, 2008). Additionally, by
inference, the greater the distance from the neutral axis, the greater the magnitude of stresses
to which the bone has adapted over time (Nordin & Frankel, 2001; Ruff, 2008; Wainwright,
1988). In long bones and in some other elements, especially tubular-shaped bones (e.g.,
metacarpals), the cross-sectional area and the manner in which the bone is distributed about
its long axis reflect mechanical/functional behavior.
Bending a ruler is a good analogy for demonstrating the principles associated with beam
analysis. If one attempts to bend a ruler against its narrow edges, there is little or no give. In
contrast, if forces are applied virtually anywhere along its flat surface, especially toward the
middle, the ruler bends readily. From a mechanical perspective, the small amount of give
when applying bending forces to the narrow edges occurs because the materials in this axis
are distributed relatively far from a central, neutral axis. Thus, in this plane of bending, the
ruler has a great deal of resistance. Conversely, the ease of deformation when applying
bending forces to the flat surface is made possible by the lack of material far from the neutral
axis; thus there is very little resistance to bending of material when the ruler is subjected to
bending forces in this direction. A ruler is structured so as to resist bending from one direction
only. Given the tubular shape of long bones, they are able to efficiently withstand mechanical
demands associated with bending and torsion from various directions, although as noted
later, some bone sections are more adapted to loads applied from particular directions.
Beam analysis involves the measurement of geometric properties from cross-sections taken
perpendicular to the long axis of a skeletal element. As demonstrated by the ruler analogy, these

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
218 Activity patterns: 2. Structural adaptation

(a)

NEUTRAL AXIS

(b)

NEUT
RAL
AXIS

Figure 6.2 Cross-section of a bone undergoing bending (a) and torsion (b), and showing
stress distribution around the neutral plane and axis, respectively. Note that the magnitude
of forces (indicated by heavier arrows) is greatest at the periphery of the bone and least
nearest the neutral plane or axis. (From Nordin & Frankel, 2001; reproduced with
permission of authors and Williams & Wilkins.)

properties are based on both the amount and the distribution of bone tissue in the cross-section.
As such, they are direct measures of the aforementioned rigidity and strength of the bone cross-
section. Therefore, the properties measured in archaeological human bones should represent a
measure of the cumulative forces operating on the skeletons of individuals during their lifetimes.
Cross-sectional geometric properties measure the amount and the distribution of skeletal
tissue in a section. For measuring rigidity, these properties include section “areas” and
“second moments of area” (or “area moments of inertia”) (Figure 6.3). Areas include total
subperiosteal area (TA), endosteal or medullary area (MA), and cortical area (CA). Measure-
ment of mediolateral (ml) and anteroposterior (ap) breadths of long bone diaphyses can be
used to calculate areas using formulae for a cylinder (Ruff & Jones, 1981):
TA ¼ π(Tap/2)(Tml/2)
MA ¼ π(Map/2)(Mml/2)
CA ¼ TA – MA

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 219

Medial-lateral
bending

Ix

Anterior-posterior
bending
Iy

Femur midshaft

Figure 6.3 Cross-section of the femur midshaft and associated geometric properties.
Second moments of area about the x- and y-axes (Ix and Iy) are associated, respectively,
with anterioposterior bending and mediolateral bending of the femoral midshaft.

where T is the total diameter of the cross-section and M is the medullary diameter. However,
this technique only works in bone regions with relatively regular, circular cross-sections, and
even there it is subject to error compared to more direct measurements (O’Neill & Ruff, 2004).
Cortical area is a measure of the amount of cortical bone in a cross-section and is also an
indicator of strength of the long bone diaphysis under pure axial loading (loading that is
simultaneously applied to both ends of the bone along its long axis). The percentage of
cortical area (PCCA) where PCCA ¼ (CA/TA)  100 is an alternative indication of the amount
of compact cortical bone, but it has no direct mechanical significance (Ruff, 2008). Cortical
area and percentage of cortical area provide very different representations of bone mass
owing to the fact that the latter, but not the former, measures cortical bone relative to total
subperiostial area (Ruff, 2008). Total subperiostial area and medullary area are measurements
of the two major surfaces of the bone cortex, including the outer periosteal and inner
endosteal surfaces, respectively (Ruff, 1991). Expansion in total subperiostial area and medul-
lary area indicates a greater distribution of skeletal tissue further from the neutral axis of
the bone.
Bone areas, especially cortical area, are proportional to strength in compression and tension
when the forces are applied non-eccentrically along the central longitudinal axis of the bone
diaphysis (axial loading). Because long bones are curved and are affected by muscular forces
applied off-center and at angles to the longitudinal axis, pure axial loadings in either
compression or tension are rare; most forces involving long bone diaphyses are eccentrically

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
220 Activity patterns: 2. Structural adaptation

applied, thus create bending and/or torsion. Second moments of area have been shown
generally to be more accurate indicators of bone rigidity and mechanical function than areas
alone. For example, analysis of second moments of area in second metacarpals from
the dominant hand in a large sample of urban Americans (n¼992) reveals that in both left-
handed and right-handed individuals the metacarpals in the dominant hand have signifi-
cantly greater second moments of area than metacarpals in the nondominant hand (Roy et al.,
1994). Increased bone strength is not the result in this case of greater cortical thickness.
This finding underscores the point that cortical thickness by itself is not an appropriate
indicator of functional/mechanical demand (Ruff, 1992; Ruff & Larsen, 1990; Trinkaus &
Ruff, 2012).
Second moments of area are geometric properties that are used to measure bending and
torsional rigidity. Bending rigidity, values of which are called “I” with a subscript that
references the specific axis running through the cross section (e.g., Ix), is calculated in relation
to the neutral axis. Typically, diaphyseal geometry is assessed at specific percentages of length
of the bone. For example, the 50% section for the femur refers to the diaphyseal geometry at
the midshaft (Figure 6.3).
The general formula for calculating I in relation to a particular axis is:
X
I¼ ai d i 2
i

where ai is the unit area and di is the perpendicular distance from the center of the unit to the
neutral axis. Ix refers to the bending rigidity in the anteroposterior plane and Iy to bending
rigidity in the mediolateral plane (Ruff, 2008). Other values of I expressing the maximum and
minimum bending rigidity in a cross-section are referred to as Imax and Imin, respectively,
where Imax measures the maximum rigidity in resistance of bone to bending and Imin measures
the minimum rigidity in resistance of bone to bending. Torsional rigidity is calculated with
reference to the neutral center or “centroid” of the cross-section and is called the polar second
moment of area or “J.” J is equal to the sum of the values of Imax and Imin (or Ix and Iy), which
are always perpendicular to each other (Ruff, 2008). J is proportional to torsional rigidity and
the average bending rigidity in any two perpendicular planes. Therefore, it represents a good
measure of overall rigidity (Ruff, 2008).
Values of bending rigidity (I) and second moment of area (J) are calculated as products of
very small unit areas in the cross-section and squared distances of the unit areas relative to
the neutral axis (for values of I) or the neutral center of the section (for values of J). Therefore,
second moments of area are presented as linear dimensions to the fourth power. Because of
variability in body size or length of long bones in comparison with different population
samples or temporal series within a particular setting, properties should be size-standardized
when making comparisons between or within groups (Ruff, 2008). Earlier work suggested that
for the femur, dividing areas by length2 and second moments of area by length4 was an
appropriate means of size standardizing (Ruff, 1984; Ruff & Larsen, 1990). More recent
analyses utilizing additional and more extensive samples of extinct and recent humans
suggest that more appropriate powers for size standardization of the femur are bone length3
and bone length5.33 for areas and second moments of area, respectively (Ruff et al., 1993).
Lastly, development of estimates of body mass via femoral head breadth provides another
method of size standardization (Trinkaus & Ruff, 2012).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 221

Cross-sectional geometric properties used to measure bending and torsional strengths are
slightly different from those used to measure rigidities. Maximum stresses occur on the outer
surface of the bone cross-section. Thus, strength cross-sectional properties (or section moduli)
are calculated by dividing the second moments of area described for rigidity by the distance
from the bone surface to the neutral axis. Section modulus values can also be estimated by
taking the rigidity values to the power of 0.73 (Ruff, 2008). These section moduli properties for
measuring strength are designated as values of “Z” with the subscript describing the plane
(e.g., Zx), with torsional strength expressed as the value Zp, representing the sum of values
perpendicular to each other and representing a measure of torsional loading. Regardless of
which are analyzed, rigidities or strengths, both variables present a record for interpreting the
same basic components of mechanical loading, with each reflecting relative patterns of
mechanical demand in past populations. Indeed, some refer to strength and rigidity under
the general term of strength, even though technically they relate to the specific meanings
described here.
Cross-sectional geometric analysis was first applied to samples of greater than a single bone
in a variety of settings involving both human and non-human primates (Jungers & Minns,
1979; Kimura, 1971, 1974; Klenerman et al., 1967; Lovejoy et al., 1976; Martin & Atkinson,
1977; Miller & Piotrowski, 1977; Minns et al., 1975; Piziali et al., 1976). These studies were
generally limited to fewer than 10 individuals because of the lengthy and tedious process
involved in manually calculating geometric properties. Specifically, two problems with
calculating section properties are the determination of endosteal and periosteal boundaries
and the mathematical integration of areas that are required for the calculations (Ruff, 2008).
The development of automated protocols for computer analysis of large numbers of cross-
sections (Nagurka & Hayes, 1980) and new technologies have made it possible to carry out
studies involving more extensive samples. These developments have fostered a more compre-
hensive understanding of variability both within and between populations from archaeo-
logical and paleontological contexts (Bridges, 1989b, 1991a, 1995; Bridges, Blitz et al., 2000;
Brock, 1985; Brock & Ruff, 1988; Churchill, 1994; Churchill & Smith, 2000; Holt, 2003; Holt &
Formicola, 2008; Kimura & Takahashi, 1982; Larsen & Ruff, 1991, 1994; Larsen, Ruff et al.,
1995; Larsen et al., 2007, 2013; Lieverse et al., 2011; Maggiano et al., 2008; Rhodes & Knüsel,
2005; Robbins et al., 1989; Ruff, 1987, 1989, 1991, 1994b, 2010a, 2010b; Ruff & Hayes, 1982,
1983a, 1983b; Ruff & Larsen, 1990, 2001; Ruff et al., 1984, 1991, 1999; Stock et al., 2011;
Stock & Pfeiffer, 2001; Sumner et al., 1985; Trinkaus et al., 1994; Van Gerven et al., 1985).
The method of geometric analysis involves the preparation of an image of a section made
perpendicular to the longitudinal axis of the bone. This image is obtained through several
alternative means, from existing breaks on long bone diaphyses (Lovejoy & Trinkaus, 1980),
direct cutting (Larsen & Ruff, 1994; Ruff & Hayes, 1983a, 1983b; Ruff et al., 1984), or
noninvasive imaging, especially computed tomography (CT) (Brock & Ruff, 1988; Bridges,
1989b; Larsen et al., 2013; Larsen, Ruff et al., 1995; Ruff, 1987, 1999), or a combination of
methods (Larsen et al., 2013; Ruff & Larsen, 2001). Other useful noninvasive techniques
include biplanar radiography (Biknevicius & Ruff, 1992; Fresia et al., 1990; O’Neill & Ruff,
2004; Runestad et al., 1993; Trinkaus & Ruff, 1989) and photon absorptiometry (Martin &
Burr, 1984; Van Gerven et al., 1985; see discussion of imaging techniques by Ruff, 1989,
2008; Ruff & Leo, 1986). Noninvasive imaging is useful in situations where cutting of bone
specimens is not possible (e.g., fossil hominins). Unlike invasive cutting, the advantage of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
222 Activity patterns: 2. Structural adaptation

noninvasive analysis is that section properties can be determined directly from the images
(e.g., CT scan) using a variety of software. More importantly, noninvasive methods leave the
bone intact.
Accuracy of geometric analysis is dependent upon the availability of fully intact periosteal
and endosteal section contours. In addition to well-preserved periosteal and endosteal sur-
faces, the accuracy of analyses – including size (length) standardization as well as precise
location of sections – requires the presence of largely intact ends of long bones. Therefore,
even though large numbers of skeletons may be present in an archaeological series, if
differential preservation exists, then only subsamples composed of well-preserved specimens
can be analyzed (Larsen et al., 2013; Ruff, 2010a, 2010b; Ruff et al., 1984).
There are various ways of obtaining landmark data for section analysis. Once the section
image is obtained, a magnified photographic slide image can be projected onto a digitizer
screen, and the periosteal and endosteal borders then manually digitized or scanned using an
optical scanner and directly input to a desktop or laptop computer. Automated computer
programs using engineering principles based on standard algorithms are widely available
(Nagurka & Hayes, 1980; Sylvester et al., 2010) for calculating section properties.
In summary, studies of cross-sectional geometric properties address a range of long-
standing issues in anthropology dealing with activity and physical behavior, especially in
regard to subsistence strategies and tool and weapon use, the relationship of sex differences to
dietary and subsistence adaptation, and variability in skeletal growth and development in
response to environmental and dietary change.

6.2.1 Specific lifeway patterns


Perspectives on physical activity are revealed by the study of patterns of biomechanical
variation within the context of specific places and times. This comparative approach reveals
striking patterns that characterize different kinds of human adaptive strategies.

American Great Basin hunter-gatherers


Analysis of cross-sectional geometry in the American Great Basin reveals a wealth of findings
on common features and variation for reconstructing workload and activity from a region of
North America having a rich archaeological and ecological record (Grayson, 2011; Kelly,
2001; Simms, 2010; Thomas, 1988) and supplemented by accounts written by nineteenth-
century pioneers and explorers (Simpson, 1983) and by twentieth-century ethnographers
(Fowler, 1992; Steward, 1938; Wheat, 1967). The sudden availability of three series of human
remains in the region owing to erosion and exposure of hundreds of archaeological sites
following record flooding in the 1980s in the western Great Basin (Stillwater Marsh), northern
Great Basin (Malheur Lake), and eastern Great Basin (Great Salt Lake) provided the opportun-
ity to reconstruct and evaluate key aspects of adaptation in an unprecedented fashion,
especially with regard to testing alternative hypotheses about use of wetlands resources,
involving either low levels of mobility (limnosedentary) or high levels of mobility (limno-
mobile) (see Chapter 5).
Analysis of cross-sectional geometry in femora (50% and 80% sections) and humeri (35%
section) from prehistoric Great Basin skeletal samples reveals important skeletal adaptations

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 223

(a) Cross-Sectional Areas


Midshaft Femur Subtrochanteric Femur Mid-Distal Humerus
400 400 400
Males
Females
350 350 350
TA
TA
300 300 300 TA

250 250 250


Areas

200 200 200

CA CA
150 150 150 CA

100 100 100


PCCA PCCA PCCA

50 50 50
Georgia Ag.

Georgia Ag.

Georgia Ag.
Pecos

Pecos
Georgia Preag.

Georgia Preag.

Georgia Preag.
Stillwater

Stillwater

Stillwater
Gr. Pl. Coal
Gr. Pl. Precoal

Figure 6.4a Comparisons of cross-sectional areas – total subperiosteal area (TA), cortical
area (CA), and percentage of cortical area (PCCA) – for three long bone sections. Note that
for each graph the foragers are at the left (Stillwater at far left) and farmers are at the right.
Areas are standardized over bone length2 and multiplied by 105. (From Larsen, Ruff et al.,
1995; reproduced with permission of American Museum of Natural History.) Cultural units:
Stillwater (1300 BC–AD 1300); Georgia Preagricultural (500 BC–AD 1150); Great Plains
Precoalescent (AD 400–1600); Pecos Pueblo (AD 1300–1650); Great Plains Coalescent (AD
1600–1850); and Georgia Agricultural (AD 1150–1550). Filled squares, males; open
squares, females.

in upper and lower limbs in a harsh desert setting (Larsen, Craig et al., 1995, Larsen, Ruff
et al., 1995, 2008; Ruff, 1999, 2010a, 2010b). Comparisons of percentage of cortical area
(PCCA), a relative measure of bone mass, with other North American archaeological series
reveal that the Stillwater group (but not Malheur or Great Salt Lake) is on the low end of a
continuum (Figure 6.4a). In sharp contrast, total subperiosteal area and J are remarkably
high in the Great Basin generally. These findings indicate that there is a relatively low
amount of femoral cortical bone for the Stillwater population but the bone that is present is
distributed far from the neutral axes relative to that for many other populations, indicating
very high bending and torsional strength. Humeral cortical area is generally reduced in all
three Great Basin samples, possibly a reflection of dietary insufficiency. As a whole, the Great
Basin samples have relatively low percentage of cortical area, that is, thin cortices, compared
to Native Americans from other regions (Ruff, 1999, 2010a), which may reflect the combined
effects of systemic stress (relatively poor diet) and a high level of physical activity.
Females and males from these various North American settings show somewhat different
patterns of skeletal morphology (Larsen, Craig et al., 1995, Larsen, Ruff et al., 1995, Larsen
et al., 2008; Ruff, 1999, 2010a). In males, torsional rigidity (J) in the femoral midshaft closely
parallels the subsistence strategy – hunter-gatherers tend to be high (e.g., Great Basin, Georgia
preagricultural), and conversely, agriculturalists tend to be low (e.g., Pecos Pueblo). This trend

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
224 Activity patterns: 2. Structural adaptation

(b) Polar Second Moment of Area


Midshaft Femur Subtrochanteric Femur Mid-Distal Humerus
200 200 200
190 190 190
180 180 180 Males

170 170 170 Females

160 160 160


150 150 150
140 140 140
J

130 130 130


120 120 120
110 110 110
100 100 100
90 90 90
80 Georgia Ag. 80 80

Georgia Ag.

Georgia Ag.
Georgia Preag.

Georgia Preag.

Georgia Preag.
Pecos

Pecos
Stillwater

Stillwater

Stillwater
Gr. Pl. Coal
Gr. Pl. Precoal

Figure 6.4b Comparison of polar second moment of area in torsional analysis (J )


in Stillwater femora and humeri with other skeletal series. Cultural units: Stillwater
(1300 BC–AD 1300); Georgia Preagricultural (500 BC–AD 1150); Great Plains Precoalescent
(AD 400–1600); Pecos Pueblo (AD 1300–1650); Great Plains Coalescent (AD 1600–1850);
and Georgia Agricultural (AD 1150–1550). Filled squares, males; open squares, females.
(Data from Ruff 1994b, 1999; Ruff & Larsen, 2001.)

follows the rationale for a decline in mechanical loading of the femur in agriculturalists in
comparison with hunter-gatherers, as a more sedentary lifeway associated with farming
emerged and intensified. This is not to say that foragers are subjected to high levels of
mechanical loading and agriculturalists are not. Rather, these biomechanical studies suggest
tendencies in patterns of mechanical loading from high levels in foragers to relatively lower
levels in agriculturalists, likely associated with different degrees of mobility.
Torsional loading bears no apparent relationship to subsistence strategy in females in these
North American groups. Rather, female torsional loading corresponds with degree of rugged-
ness of terrain – mountainous populations (Great Basin, Pecos Pueblo) have the highest values
of J, coastal populations have the lowest values (Georgia coast), and Great Plains populations
are intermediate (Figure 6.5).
Humeri show a somewhat different pattern of area and second moments of area than
femora. Especially striking are the low values of humeral cortical area and second moment of
area (J), notably in the Stillwater Great Basin series (Larsen, Ruff et al., 1995; Ruff, 2010a).
This finding is consistent with the hypothesis that mechanical loading is localized in the
skeleton (Lanyon & Skerry, 2001; Macdonald et al., 2009; Peck and Stout, 2009).
Thus, the generally high level of robusticity in Great Basin adults – especially in femora – is
consistent with findings based on osteoarthritis prevalence (and supporting a limnomobile
subsistence strategy; see Chapter 5). Behaviors causing osteoarthritis and elevated robusticity
result in respective high frequencies of joint degeneration and high second moments of area.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 225

550

Relative Femoral Strength (J)


500

450

400

350
Mountains Plains Coastal
Terrain
Figure 6.5 Torsional rigidity correlates with terrain, with mountainous populations (Great
Basin and Pecos Pueblo) showing markedly higher relative femoral strength (J) than coastal
populations (Georgia Bight). (Ruff, 2008; reproduced with permission of author and John
Wiley & Sons, Inc.)

Although Great Basin foragers from all three settings show quite high levels of skeletal
robusticity, for the lower limbs, some populations from the region of the Great Basin pursued
a maize farming strategy. That is, stable carbon isotope analysis of the Great Salt Lake later
prehistoric Fremont (AD 400–1150) shows an increasing reliance on maize (Coltrain &
Stafford, 1999; and see Chapter 8). Interestingly, individuals having a strong isotopic signa-
ture for maize consumption have generally lower skeletal robusticity (J) for both femora and
humeri compared to individuals who have a weak isotopic signature for maize consumption,
with the pattern more pronounced for males (Ruff, 1999) (Figure 6.6). These findings show a
clear link between dietary regime, behavior, and their combined influence on skeletal
morphology.

Subarctic hunter-gatherers
Populations inhabiting the Aleutian Islands, Alaska, from the mid-eighteenth century to the
present, also led physically demanding lifestyles. An especially demanding task was kayaking
on the open ocean (Laughlin et al., 1991, 1992; and see Chapter 5). This activity underlies the
extreme external robusticity of Aleut humeri (Hrdlička, 1945) and high levels of osteoarthritis
and musculoskeletal stress markers in these populations (Hawkey & Street, 1992; and see
Chapter 5). Comparisons of Aleuts with other populations in humerus torsional rigidity reveal
that these subarctic peoples surpass values derived from a range of modern groups (Churchill,
1994; Ruff et al., 1993) (Table 6.1). These elevated levels of bone rigidity in adult humeri
reflect intensive mechanical loading of the upper limb, especially in use of the arms in
paddling of watercraft on the open ocean. Moreover, the findings are consistent with the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
226 Activity patterns: 2. Structural adaptation

(a)
800
Females
Femoral Size-Standardized J 700 Males

600

500

400

300

200

100

0
-20 -18 -16 -14 -12 -10
d13C (ppm)

(b)
1100

Females
1000
Males
Humeral Size-Standardized J

900

800

700

600

500

400

300
-20 -18 -16 -14 -12 -10
d13C (ppm)

Figure 6.6 Femoral (a) and humeral (b) relative strength (J) are plotted relative to stable
carbon isotope ratios. Stable carbon ratios reflecting maize-rich diets (less negative
δ13C values) are associated with lower femoral and humeral robusticity. (Ruff, 1999;
reproduced with permission of author and University of Utah Press.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 227

Table 6.1 Humerus polar moments of area (35%, 50%, and 65% sections) in selected fossil
and recent human adults (Adapted from Churchill, 1994: Tables 16 and 17)
Females Males

35% 50% 65% 35% 50% 65%


Group
Euroamerican 79.9 89.7 96.8 124.3 145.5 149.3
(n) (19) (19) (19) (25) (25) (25)
Afroamerican 101.9 113.5 117.8 151.4 169.6 178.8
(n) (25) (24) (25) (25) (25) (25)
Aleut 141.2 184.7 210.0 198.6 265.6 316.5
(n) (21) (22) (22) (25) (25) (25)
Amerindian 102.3 121.4 128.3 102.1 118.5 124.8
(n) (20) (20) (20) (20) (20) (20)
Peruvians 66.1 86.5 95.6 127.6 139.7 147.3
(n) (1) (1) (1) (3) (3) (3)
Late Upper Paleolithic 125.2 155.0 172.1 154.3 183.8 183.0
(n) (6) (6) (5) (9) (9) (9)
Early Upper Paleolithic 82.0 132.8 120.0 144.0 128.8 192.7
(n) (4) (3) (3) (6) (6) (5)
Skhul/Qafzeh 101.3 79.5 95.9 92.2 82.9 123.5
(n) (1) (2) (2) (1) (2) (1)
Archaic humans 125.2 133.3 151.0 166.7 184.2 194.9
(n) (4) (4) (4) (7) (6) (6)

analysis of cross-sectional geometry of rowers whereby rowers tend to have high levels of
robusticity, but especially in populations rowing in open ocean settings (Ruff, 2006; Weiss,
2003).

6.2.2 Bilateral asymmetry in humeral loading


Microwear orientation on stone tools (Toth, 1985) and on anterior teeth of early hominins
(Bermúdez de Castro et al., 1988; Frayer et al., 2012) indicates that humans have had right-
sided dominance of the upper limb for much of the Pleistocene. These lines of evidence
represent indirect means of assessing and interpreting upper limb use. A more comprehensive
understanding is provided by the study of structural bilateral asymmetry of limb bones, a
subject of interest to anthropologists and anatomists since the mid-nineteenth century
(Dangerfield, 1994; Fresia et al., 1990; Schaeffer, 1928; Shaw & Stock, 2009a; Stirland,
1993). Asymmetry is of particular interest to biological anthropologists because some degree
of left–right-side differences has been documented in a variety of human populations and has
played a key role in discussions of genetic versus functional explanations of bone size and
morphology.
All human populations express higher frequencies of right than left dominance for external
measurements of the humerus, including length, thus indicating a probable genetic

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
228 Activity patterns: 2. Structural adaptation

component. There is certainly a strong genetic determinant for handedness, resulting in


asymmetry in the comparison of left and right limbs. However, the differing patterns of
asymmetry in relation to specific levels and types of activity in various skeletal samples
investigated by biological anthropologists argue for a functional interpretation for diaphyseal
long bone morphology (see discussions in Auerbach & Ruff, 2006; Martin et al., 1998; Sládek
et al., 2007; Trinkaus et al., 1994). Moreover, ontogenetic analysis of patterns of asymmetry
shows that left–right differences in size begin only in later childhood, strongly supporting a
developmental/functional model (Blackburn, 2011).
A wide variety of behavioral, lifestyle, and occupational evidence shows that use of the
upper limbs is variable within and between populations. Some activities involving the upper
limb involve both limbs equally, and thus, equal levels of mechanical loading (e.g., pounding
of maize in a wooden mortar and pestle; Bridges, 1991a). Some activities, however, focus
heavily on one upper limb with associated unilateral loading of the dominant limb, such as
professional tennis players (Haapasalo et al., 1996; Trinkaus et al., 1994) or the use of the
bow-and-arrow versus spear throwing or thrusting in hunting (Schmitt et al., 2003; Stock &
Pfeiffer, 2004). With regard to spear throwing or thrusting, for example, experimental
evidence indicates that extraordinary levels of force are required for both distance and
accuracy (Schmitt et al., 2003). Thus, the investigation of bilateral asymmetry in humeri from
archaeological series is an important means of reconstructing and interpreting activity in
the past.
Comparisons of structural properties – areas (CA, MA, TA) and second moments of area
(Ix, Iy, J) – from precontact preagricultural, precontact agricultural, early contact, and late
contact humeri (35% section) from the Georgia Bight populations provides important
perspective on behavioral shifts (Fresia et al., 1990). These comparisons reveal a pattern
of right-dominant bilateral asymmetry in this setting, similar to findings reported by other
researchers (Ben-Itzhak et al., 1988; Borgognini Tarli & Repetto, 1986; Bridges, 1989b,
1991b; Constandse-Westermann & Newell, 1989, 1990; Hrdlička, 1932; Ledger et al., 2000;
Rhodes & Knüsel, 2005; Ruff & Hayes, 1983b; Ruff & Jones, 1981; Stirland, 1993, 2002;
Trinkaus et al., 1994). Specific changes in pattern of asymmetry indicate shifts in the
manner in which the upper limbs were used from the earliest to the latest period. Analysis
of left–right torsional rigidity (J) of the mid-distal humerus (35% section) shows a general
decrease in asymmetry through the early contact period; asymmetry then increases in the
late contact period.
Comparisons of adult males and females reveal a greater decline in asymmetry in females
than in males in the transition from foraging to farming prior to European contact. This
suggests that change in the use of the upper limbs was more profound for females than for
males in the shift to agriculture. The relatively greater change in upper limb biomechanical
loading in females is consistent with the notion that subsistence activities changed more in
women than in men: women were engaged in labors of food preparation (e.g., maize
pounding), whereas men continued to follow a lifestyle similar to that of the earlier prehistoric
period (e.g., hunting). The asymmetry between left and right sides declines in the precontact
agricultural women to virtually nil. This finding adds confirmation to Bridges’ (1991b)
hypothesis that American Southeast females had relatively equal use of left and right arms
in maize pounding and grinding and activities that generally required the simultaneous use of
both arms.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 229

Comparisons of sexual dimorphism – the percentage of side differences in J between males


and females – shows a marked decline in the temporal span, culminating with the least amount
of sexual dimorphism in the early and late contact periods. This finding lends strong support
for the hypothesis that contact-era men and women engaged in physical activities involving
comparable loading modes. The trend for reduction in sexual dimorphism in torsional rigidity
and other structural properties suggests that division of labor between sexes reduced dramat-
ically in intensive agriculturalists in the contact period. Historical accounts (Hann, 1986)
indicate that some men undertook tasks normally performed by women, especially with regard
to activities involving the use of the arms. This convergence of male and female behaviors
would contribute to a reduction in sexual dimorphism in upper limb bilateral asymmetry. Other
historical accounts suggest that male and female activities were different (Swanton, 1946). The
possibility remains, therefore, that the increasing similarity of mechanical loading is due to
other behavioral changes that coincidentally resulted in comparable cross-sectional geometry.
The reduction in bilateral asymmetry in the Georgia Bight is similar to the pattern observed
in comparison with prehistoric hunter-gatherers and agriculturalists from the Pickwick Basin,
Alabama (Bridges, 1989b) and other North American settings where maize agriculture was
adopted (Ogilvie & Hilton, 2011). Comparison of external diaphyseal dimensions (data are not
available for left–right asymmetry of geometric properties) reveals that females decline in
asymmetry, which may reflect the use of both arms in preparation of maize (Bridges, 1989b).
Males also show a reduction in asymmetry of left and right humeri, but it is slight in
comparison with that of females. Thus, over the course of time, female and male activities
in both southeastern United States settings probably became more similar, especially with
regard to those activities involving the use of the upper limbs.
The results from Georgia and Alabama are different, however, from the study of a temporal
series dating the transition from the Late Eneolithic (2900–2000 BC) to the Early Bronze Age
(2200 BC–1700 BC) in central Europe (Sládek et al., 2007). In this setting, there was no change
in biomechanical asymmetry and, therefore, no clear change in differential use of the left
versus right upper limb. In this setting, there is clear archaeological evidence of economic and
social changes. However, both periods involved intensive agricultural production with very
little hunting. This suggests that the dietary changes identified in the North American series
and associated work involving considerable changes in food production is an important factor
in understanding why a similar economic transition affected use of the upper limbs in one
setting (North American) but not another (central Europe). These investigations also point to
the crucial importance of understanding local variation in economic, social, and dietary
factors in understanding biomechanical variation.
Bilateral asymmetry of cross-sectional geometric properties of the humerus in modern
human populations shows a wide range of variation (Ruff, 1992; Shaw & Stock, 2009a;
Trinkaus et al., 1994). Several modern samples have moderately high levels of asymmetry
(5%–14%). Professional tennis players have extraordinarily high levels of asymmetry (28%–
57%), a pattern which is similar to Pleistocene late archaic hominins (24%–57%) (Trinkaus
et al., 1994). The high level of asymmetry in professional tennis players and the overall
variability documented in human samples attest to the potential for dramatic change in
diaphyseal morphology, depending upon the nature of upper limb use and function and
especially the mechanical loading of the left versus the right arms (see discussions in
Ruff et al., 1993; Shaw & Stock, 2009a; Trinkaus et al., 1994).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
230 Activity patterns: 2. Structural adaptation

Variability between status groups in bilateral asymmetry may reveal behavioral differences
within ranked societies. Degree of left–right dominance in external measurements of Meso-
lithic limb bones from Western Europe shows some difference in work and activity between
upper- and lower-status groups (Constandse-Westermann & Newell, 1989). High-status
females have reduced right dominance when compared with low-status females.
Constandse-Westermann and Newell (1989) contend that the greater upper limb lateralization
in low-status women reflects their heavier work demands than those of high-status women.
Higher-status males possess greater right dominance than lower-status males. Although
reasons for the contrasting pattern in males and females are unclear, they speculate that in
order for males to achieve high status, they were required to perform more demanding and
more differentiated tasks.
Analysis of individuals engaged in highly specific, highly demanding activity that required
the use of a dominant arm provides important perspective on the interrelationship between
form, function, and behavioral inference. Examination of cross-sectional geometry of the
humeri of battle victims in Medieval contexts from Towton (see Chapter 4) and Fishergate,
England, shows striking differences in diaphyseal robusticity and shape of the left and right
humeri (Rhodes & Knüsel, 2005). Comparison of blade-injured and non-blade-injured males
reveals distinctive patterns of variation in areas and second moments of area (CA, TA, J, and
Ix/Iy). Blade-injured males display consistently higher values for these properties compared to
non-blade-injured males from the same period. In particular, values of J, a good overall
indicator of resistance to bending/torsion, are remarkably high in the Towton battle victims.
With regard to diaphyseal shape asymmetry, values of Ix/Iy present distinctive left–right
differences in the Towton series. This pattern of bilateral shape asymmetry represents signifi-
cant differences in loading, movement patterns, and use of the arms generally, in these
individuals. These right–left side differences reflect years of training and the adaptation of
bone to profoundly demanding activities. Rhodes and Knüsel (2005) present the case that the
upper left limb in these individuals was adapted to activities that increase bone deposition,
especially in the mediolateral dimension, such as movement patterns involving abduction,
adduction, and rotation of the arm, possibly using a longbow. In contrast, Fishergate males
show right-dominance in cross-sectional properties, suggesting long-term training in use of
the right arm, which is likely related to use of a right-handed unimanual weapon, such as a
sword or lance.
While the bioarchaeological record shows highly specific patterns of humeral strength (or
rigidity) that reflect local adaptations, there is a general pattern of reduction in upper limb
asymmetry over the course of human evolution. In particular, early hominins are highly
asymmetrical, Holocene foragers are moderately asymmetrical, and Holocene farmers and
modern humans having the least degree of asymmetry in the upper limb (Figure 6.7).
Asymmetry of the lower limb is poorly known in archaeological or other population
settings. Lateralization of lower limb bones has been studied from California (Ruff & Jones,
1981), Pecos Pueblo (Ruff & Hayes, 1983b), and Mesolithic Western European (Constandse-
Westermann & Newell, 1989, 1990). These investigations reveal that the left side tends to have
slight size dominance and greater mechanical strength, especially in anteroposterior bending
forces. Generally, lower limb asymmetry in structure and overall size is either considerably
less than upper limb asymmetry or shows no consistent pattern (Auerbach & Ruff, 2006;
Borgognini Tarli & Repetto, 1986; Constandse-Westermann & Newell, 1989, 1990;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 231

40

Humeral Bilateral Asymmetry in Z p (%)


30

20

10

0
Georgia Mission

Jomon

Euroamerican

Calif. Amerindian

SW Amerindian

Aleut

Late Up. Paleolithic

Neanderthal

Early Up. Paleolithic

Tennis Players
Figure 6.7 Humeral bilateral asymmetry in Zp among Holocene and modern human
populations. Paleolithic and Neanderthal populations demonstrate humeral bilateral
asymmetry that equated with modern day tennis players. This similarity in asymmetry
indicates that early hominins engaged in activities where one arm was consistently used
more than the other. In contrast, modern humans from temperate coastal (Georgia Mission)
to Arctic (Aleut) environments show significantly less humeral bilateral asymmetry. These
results indicate generally equal distribution of mechanical loading on both left and
right arms. (Ruff, 2008; reproduced with permission of author and John Wiley & Sons, Inc.)

Dangerfield, 1994; Ruff & Hayes, 1983b). The differences in asymmetry in upper and lower
limbs in humans reflect the fact that the upper limbs are used in a wide variety of non-
ambulatory activities, whereas lower limbs are primarily used in bipedal locomotion, a super-
dominant function requiring equal use of left and right sides.

6.2.3 Temporal trends in modern and pre-modern recent humans


Biomechanical approaches to the study of physical activity and behavioral change add an
important dimension to our understanding of adaptive shifts in the past.

Eastern North American transition from foraging to farming


In the Pickwick Basin of northwestern Alabama, analysis of femoral and humeral cross-
sectional geometry reveals a number of differences between earlier Archaic-period hunter-
gatherers and later Mississippian-period agriculturalists (Bridges, 1991b). Femoral cortical
area and minimum and maximum second moments of area increase in the Mississippian

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
232 Activity patterns: 2. Structural adaptation

population in comparison with the Archaic population for both adult males and females, thus
indicating greater bone rigidity in the agriculturalists than in the foragers. Structural analysis
of male humeri shows that the two temporal series are virtually indistinguishable. Thus, for
males, activity levels increased, but primarily in relation to lower limb (ambulatory) functions.
In females, there are significant increases in humeral and femoral rigidity. Bridges suggests
that the increase in bone rigidity in female femora and humeri reflects a relatively greater
range of activity changes in them than in males. Thus, as revealed by the osteoarthritis
analysis, the shift to food production may have had a relatively greater impact on women
than men in this setting. Some of these activities are likely related to who is responsible for
food processing; increases in female humeral bone rigidity may be associated with maize
processing. Historic-era Southeastern native women used wooden mortars and pestles
that required very demanding physical labor involving the upper limbs. One nineteenth-
century observer equated the task of maize pounding to blacksmithing (Current-Garcia &
Hatfield, 1973).
These findings from northwestern Alabama suggest that the adoption of an agricultural
lifeway involved more strenuous physical activity than in earlier populations. The decline in
osteoarthritis prevalence for these populations suggested a decrease in mechanical loading
(see Chapter 5). Thus, these two indicators of mechanical stress seem to yield contradictory
results. Bridges (1991b) suggests that osteoarthritis and long bone geometry are due to
different types of activities. For example, citing findings from sports medicine and other
research, she notes that some normal activities – such as running – do not contribute to
osteoarthritis (Eichner, 1989; Panush & Brown, 1987). Less frequent movements that lead to
microtrauma and injury of joints may be important factors in the explanation of osteoarth-
ritis. It is virtually impossible to distinguish traumatic from daily “wear-and-tear” osteoarth-
ritis, thus making it difficult to identify specific causes where behaviors are unknown, such as
in archaeological settings. Skeletal structural change has been tied to long-term repetitive
forces (Lanyon & Rubin, 1984; Shaw et al., 1987). Thus, osteoarthritis and diaphyseal cross-
sectional geometry should not necessarily be concordant because osteoarthritis would be
expected to develop in older adults, whereas diaphyseal remodeling is a lifelong response to
mechanical stimuli (Ruff, 1992).
A similar pattern of change in cross-sectional geometry is documented in the lower Illinois
River valley, a region that saw a transition from mixed foraging/hunting and consumption of
domesticated plants (native starchy seed crops such as Chenopodium) during the Middle
Woodland (50 BC–AD 200) and Late Woodland (AD 600–1050), to the introduction of maize
in the later Late Woodland, and intensive maize production and consumption in the Mis-
sissippian period (AD 1050–1250). Analysis of skeletal morphology (cortical area, CA, and
second moment of area, J) in light of economic intensification associated with more intensive
plant production and especially increased maize production in the Mississippian period
revealed little evidence of change in males (Bridges, Blitz et al., 2000). In contrast, females
show a distinctive pattern of increase in mechanical loading, as it is represented in second
moment of area (J) and cortical area (CA), through the later part of the Late Woodland period.
In females, however, there is a decline in cross-sectional geometric properties in the Mis-
sissippian period. This pattern of change reveals that the economic shifts more profoundly
affected women than men, perhaps indicating that women were more heavily involved in
agricultural production and food preparation than men; men continued similar kinds of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 233

behaviors, such as those associated with hunting, throughout all four periods. The decline in
activities affecting especially the upper limbs of females in the Mississippian period may be
related to innovations in how maize was prepared in later prehistory in the lower Illinois River
valley. For example, innovations in ceramic technology that allowed increased boiling of
maize and other food may have resulted in reduced workload involving pounding and
processing, hence explaining the clear evidence for reduction in mechanical demand on
women’s upper limbs (Bridges, Blitz et al., 2000).
Analysis of femoral and humeral diaphyses from preagricultural and agricultural Georgia
Bight populations yields an unambiguous decline in second moments of area for both males
and females, reflecting a decline in workload in this setting (Larsen & Ruff, 1991, 1994, 2011;
Ruff & Larsen, 1990; Ruff et al., 1984) (Figure 6.8).
The results reported from the three regions of the southeastern United States – Pickwick
Basin, lower Illinois River valley, and Georgia Bight – are different, especially in comparison
with second moments of area (Larsen & Ruff, 2011) (Figure 6.8). However, the comparisons of
periods within each of the three settings described here are informative about workload and
activity and how they vary across landscapes and adaptive shifts. Comparison of temporal
trends for J for the femur and humerus reveals significant differences between the three
groups. For males, femoral midshaft rigidity increases in Alabama, shows no change in the
lower Illinois River valley, and decreases substantially in the Georgia Bight. In females,
femoral midshaft rigidity increases in Alabama, increases in the lower Illinois River valley
in the first three periods, but then decreases in maize farmers, and decreases in the Georgia
Bight. Comparisons of cross-sectional geometry in the humerus show broadly similar patterns
across the three settings. However, male humeri show no change in Alabama and the lower
Illinois River valley, and a decrease in the Georgia Bight. Female humeri show an increase in
Alabama and the lower Illinois River valley, followed by quite a substantial decrease in the
Mississippian period and a decrease in the Georgia Bight.
While these changes are somewhat complex in viewing the region as a whole, several
general trends are clear. First, the comparisons reveal different patterns of temporal change
in males and females in the foraging-to-farming transition. The degree of change is greater
in females in Alabama and the lower Illinois River valley, but greater in males in the
Georgia Bight. This suggests that the transition had a great impact on women in the
terrestrial, inland setting and on men in the marine, coastal setting. In two of the three
regions (Alabama, Georgia Bight), there was only one agricultural period (introduction of
maize), whereas in the third region (lower Illinois River valley), there were two agricultural
periods (local domestication of weedy plants; introduction of maize). The latter involved
changes in activity in both periods, but with different outcomes – an increase in activity in
the shift to exploitation of oily starches, followed by a decrease in activity subsequent to
the adoption of maize. The implications of the biomechanical analyses are that the transi-
tion had behavioral significance well before maize was adopted and intensified. Import-
antly, these findings strongly suggest that the agricultural transition was not a monolithic
event having the same outcome across the region of the American Southeast, or even
within a single region (e.g., lower Illinois River valley). In fact, factors of terrain and
ecology – markedly different when comparing inland and coastal settings – almost cer-
tainly influenced the kinds and levels of activity that affected skeletal growth and adult
diaphyseal morphology.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
234 Activity patterns: 2. Structural adaptation

(a)

Pickwick Basin
450 Foragers
Farmers
400
Femur JSTD

350

300

250

200
Males Females

Lower Illinois Valley


450
Middle Woodland Farmers
Early Late Woodland Farmers
400 Late Late Woodland Farmers
Mississippian Farmers
Femur JSTD

350

300

250

200
Males Females

Georgia Bight
600
Foragers
550
Farmers
500
Femur JSTD

450
400
350
300
250
200
Males Females

Figure 6.8 Second moments of area (JSTD) in femora (a) and humeri (b) from Lower Illinois
Valley and Georgia Bight populations decline in late prehistory. These results suggest a
decline in upper- and lower-body workload. In contrast, the Pickwick Basin populations
increase in JSTD values. This difference suggests that variation in mobility and activity
exists within subsistence strategies between periods and regions. (Larsen & Ruff, 2011;
reproduced with permission of John Wiley & Sons, Inc.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 235

(b)

Pickwick Basin
Foragers
Farmers
600

500
Humerus JSTD

400

300

200
Males Females

Lower Illinois Valley


700
Middle Woodland Farmers
Early Late Woodland Farmers
600
Late Late Woodland Farmers
Humerus JSTD

Mississippian Farmers
500

400

300

200
Males Females

Georgia Bight
800 Foragers
Farmers
700
Humerus JSTD

600

500

400

300

200
Males Females

Figure 6.8 (cont.)

The study of structural morphology in Georgia Bight limb bones has been expanded to
include the descendant, contact-era populations, including the descendant Guale population
on St. Catherines Island, Georgia (Santa Catalina de Guale), the later descendant Guale on
Amelia Island, Florida (Santa Catalina de Guale de Santa Maria), and an early contact series
preceding full missionization from Amelia Island representing Timucua (Larsen & Ruff, 1994,
2011; Ruff & Larsen, 1990, 2001). Cross-sectional geometric analysis for the femur reveals

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
236 Activity patterns: 2. Structural adaptation

(a)

600
Males
Femoral Midshaft Bending/Torsional Robusticity (JSTD) Females

500

400

300

200

100

0
EPG LPG EMG LMG EMT
Group

Figure 6.9 Torsional robusticity (JSTD) decreases in femoral (a) and humeral (b) midshafts
from precontact to postcontact populations on Amelia Island: EPG (Early Preagricultural
Guale), LPG (Late Preagricultural Guale), EMG (Early Mission Guale), LMG (Late Mission
Guale), and EMT (Early Mission Timucua). (Ruff & Larsen, 2001; reproduced with
permission of University Press of Florida.)

that beginning with the early contact period on St. Catherines Island, there is a reversal of the
decrease in bone rigidity that had been documented in the late prehistoric populations from
the region (Figure 6.9). Geometric properties increase in the early contact period, but decline
in the late contact period in the Guale series. The very early Timucua have values of J more
like those of the precontact agricultural Guale from St. Catherines Island for the femur.
In males, humeral second moments of area increase successively in the early and late
contact periods. In females, there is a continued decline in humeral values in the early contact
period, but this reverses in the late contact period. Thus, both females and males in the late
contact period experience a marked increase in humeral rigidity. In general, the changes
observed in bone rigidity in the contact period indicate that native populations experienced
increases in mechanical loads, probably due to increased manual labor and physical demands
placed on them by the Spanish (see Chapter 5). The early contact-era Timucua, however, are
more similar in these values compared to later contact populations. This finding is consistent

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 237

(b)

Males
800 Females
Mid-Distal Humeral Bending/Torsional Robusticity (JSTD)

600

400

200

0
EPG LPG EMG LMG EMT
Group

Figure 6.9 (cont.)

with the notion that the earliest contact populations that preceded full missionization and its
associated exploitation of native labor would be more similar to precontact populations than
to later contact populations in workload and lifestyle generally (Ruff & Larsen, 2001).
Because nutritional quality also influences skeletal size to a degree (see Chapter 2), the
possibility remains that the mechanical environment may not be the sole factor that explains
the structural modifications documented in the Georgia Bight. Standardized values of cortical
area show very little change in the comparison of the four periods, thus suggesting that bone
mass remains essentially unchanged through time. In contrast, the distribution of bone tissue
alters dramatically, which is consistent with skeletal adaptations to localized (mechanical)
factors rather than systemic (nutritional) stress (Larsen & Ruff, 1994; Ruff, 1999; Ruff &
Larsen, 2001; Ruff et al., 1984).

Great Plains biomechanical adaptations


The northern Great Plains of North America was the locus of dramatic shifts in lifeway,
including the adoption of the horse for transportation, the migration of ancestral Arikara from
the central to northern Plains, and the replacement of native populations already in place by

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
238 Activity patterns: 2. Structural adaptation

new arrivals prior to the European invasion. With the arrival of Europeans in the seventeenth
century came increased reliance on domesticated plants (maize) and farming overall, intensi-
fied long-distance hunting of bison, and intensified trade and warfare between native
populations and Euroamericans (Blakeslee, 1994; Owsley & Jantz, 1994). Much of this record
pertains to the late prehistoric Plains Village period (post-AD 950–1600) and Historic period
(post-AD 1600) and involved dramatic lifestyle and activity changes. Biomechanical analysis
contributes important insight into these developments, sometimes rapidly occurring and
limited to the region (e.g., intensified hunting with reliance on the horse) but also in
comparison, more broadly, with North America generally (maize farming). The adoption of
the horse had a profound impact on cultures in the region. In addition to altering subsistence
practices, the adoption of the horse facilitated inter-tribal trade and increased the ability to
travel long distances. In comparing cortical area (CA) and second moment of area (J) for
earlier and later Historic populations, no differences in mechanical loading of femora were
found (Ruff, 1994b). However, there is a clear decline in sexual dimorphism in these properties
in comparing males and females, suggesting that the transition involved increasingly similar
activities between the sexes, a pattern much like settings in the American Southeast (e.g.,
Georgia Bight). With regard to specific tribes, biomechanical comparisons of late prehistoric
and historic Arikara reveal that with intensified farming, a development related to increasing
economic focus on creating surpluses of crops for trade, mechanical loading increased to some
extent but primarily in females (Wescott & Cunningham, 2006). The greater work change in
women than men suggests that women may have been the primary source of labor for
increasing crop production for trade.

Foraging to farming in the American Southwest


Structural analysis of other populations undergoing the transition to agriculture show
changes in long bone diaphyses that are broadly similar to patterns observed in the Alabama
setting studied by Bridges (1989b), which show an increase in mechanical loading. In the
American Southwest, the comparison of Archaic-period foragers from southwest Texas with
later prehistoric farmers from central New Mexico reveals significant increases in cross-
sectional geometric properties for the humerus, clearly showing an increase in mechanical
loading of the upper limb, especially in females. These findings provide additional clarity
regarding the impact of the foraging-to-farming transition on women as it relates to workload
involving the upper limb, such as in preparation of maize flour and other agriculture-related
behaviors (compare with Bridges, 1989b; Bridges, Blitz et al., 2000; Larsen & Ruff, 2011).
Analysis of a temporal series from western New Mexico representing adoption and
intensification of agricultural production – Early Village (AD 500–1150), Abandonment
(AD 1150–1300), and Aggregated Village (AD 1150–1300) – reveals a different pattern of
skeletal adaptation than in central New Mexico and western Texas. Early Village populations
focused primarily on nondomesticated plants and animals, but likely there was some farming;
Abandonment populations experienced a transition to new adaptive patterns and new cir-
cumstances, resulting in a state of adaptive “disequilibrium”; and Aggregated Village popu-
lations were sedentary maize agriculturalists living in large villages (Brock & Ruff, 1988).
Second moments of area (Imax, Imin, J) show a general increase in the Abandonment period
and either remain constant or decline in the Aggregated Village period. In males, values for
the femoral midshaft show a clear pattern of decline from earliest to latest periods. These

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 239

86
50% Section Femur
84
50% Section Tibia
Mean Cortical Percentage (%CA)
82

80

78

76

74

72

70

68
Early Mid Neanderthals MPMH EUP/MUP
Pleistocene Pleistocene
Pleistocene Homo
Figure 6.10 Temporal reduction in femoral midshaft robusticity in Homo from the Early
Pleistocene to Upper-Middle Paleolithic. (Adapted from Trinkaus & Ruff, 2012; reproduced
with permission of authors and PaleoAnthropology Society.)

findings suggest an overall decrease in mechanical demand with agricultural intensification


and sedentism during the Aggregated Village period. The ratio of Ix/Iy in the femur midshaft
shows a decline in both sexes, which also suggests a reduction in bending stresses –
particularly in the anteroposterior plane – as populations became less mobile during later
prehistory (see later). These observations are consistent with archaeological reconstructions of
increasing sedentism with the shift to agriculture in the American Southwest (Brock &
Ruff, 1988).
Behavioral reconstructions of pre-modern human populations based on structural analysis
are not usually as clearly defined as the previously discussed groups, in large part owing to the
small numbers of samples available for study. Several key analyses serve to provide a broader
understanding of pre-modern human skeletal structural variation in relation to activity and
behavior. Geometric analysis of a limited sample of Neanderthal tibiae – from Shanidar,
Amud, and La Chapelle-aux-Saints – reveals that bending and torsional rigidity is of the order
of twice that observed in modern humans from the late prehistoric Libben site, Ohio (Lovejoy
& Trinkaus, 1980). Comprehensive study of fossil hominin and early modern human groups
encompassing the evolution of the genus Homo shows increased relative bone mass (repre-
sented by percentage of cortical area) and continued high levels of activity in adulthood (Ruff
et al., 1994; Trinkaus & Ruff, 2012) (Figure 6.10). When accounting for body mass via
estimates based on femoral head size, there is little change in overall rigidity and implied
mobility throughout most of the Pleistocene (Trinkaus & Ruff, 2012). However, during the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
240 Activity patterns: 2. Structural adaptation

Late Pleistocene and early Holocene – between the Upper Paleolithic and Neolithic – there is a
significant reduction in lower limb rigidity and mobility (Holt et al., 2012; Shaw & Stock,
2013), indicating major alterations in mechanical loading in general and mobility in
particular.
Femoral head size relative to body size also remains relatively constant throughout most of
the Pleistocene (Ruff et al., 1993). These findings lend support to the hypothesis that joint
morphology is more genetically canalized or less developmentally plastic than long bone
diaphysis morphology. In adults, articular joint size and shape does not alter in response to
mechanical loading, unlike diaphyseal structure (Rafferty & Ruff, 1994; Ruff & Runestad,
1992; Ruff et al., 1991, 1994). Changes in articular loading in adults can have profound effects
on subchondral and trabecular bone structure organization underlying the joint surface
(Pauwels, 1976; Poss, 1984; Radin et al., 1982, 1984; Rafferty & Ruff, 1994; Ruff, 1992,
2008), but these changes are not manifested in external articular size. Thus, internal and
external joint structures are independent and represent contrasting expressions of the mech-
anical environment.
The lack of influence of mechanical loading on joint size has important implications for
functional studies of archaeological remains, especially where joint size differences are used
to infer mechanical differences between groups. For example, increase in femoral head size in
comparing prehistoric hunter-gatherers and agriculturalists in the Caddo region of the Ameri-
can Southeast was interpreted to reflect an increase in protein consumption and increase in
mechanical loading (Rose et al., 1984). These new findings comparing articular and diaphy-
seal structure indicate that mechanical loading as a causal factor in explaining temporal
change in joint size is highly unlikely.

Foraging to farming in the Nile Valley


Biomechanical analyses have provided essential means by which to understand the foraging-
to-farming transition in a region of the world where intensive agriculture developed. This
transition provided the context for the rise of complex societies, including the Pharaonic state
in Egypt associated with construction of pyramids and other elaborate monumental architec-
ture and the reliance on specialized laborers. As with many other areas of the world, the shift
from a lifeway based exclusively on hunting and gathering to agriculture in the Nile Valley
(Egypt and Sudanese Nubia) had a profound impact on every aspect of life, including labor
and workload. In this setting, the earliest evidence of food production in Neolithic cultures
indicates that the cultivation of barley and management of cattle, sheep, or goats began
sometime around 5000 BC (Hassan, 1988). Building on other findings for the agricultural
transition (see earlier), Stock and coworkers (2011) test the hypothesis that workload generally
declined with the adoption of agriculture in this setting. Comparisons were carried out on Late
Paleolithic (Jebel Sahaba; 9000–13 000 BC) hunter-gatherers with a succession of three
agricultural periods (Predynastic el-Badari, 5000–4000 BC; Predynastic Hierakonpolis,
4000–3000 BC; Dynastic Kerma, 2100–1500 BC), focusing on second moments of area (Ix,
Iy, Imax, Imin) and polar moment of area (J). In support of their hypothesis, there was a
pronounced decline in humeral robusticity (J) following the Late Paleolithic, but in males
only (Figure 6.11). In the subsequent periods, humeral J remains constant over time. Females,
however, have a considerably lower humeral robusticity than males in the Late Paleolithic
sample, which increases somewhat in the early Predynastic period, followed by a decline. In

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 241

250

Male
Female
200
Standardized Humerus J

150

100

50

0
Jebel Sahaba el-Badari Hierakonpolis Kerma
Site
Figure 6.11 Decline in male humeral robusticity (J) from preagricultural (Jebel Sahaba) to
agricultural (el-Badari, Hierakonpolis, and Kerma) Egyptian populations. Females show no
significant difference in humeral robusticity. (Burger & Thomas, 2011; reproduced with
permission of author and John Wiley & Sons, Inc.)

males, femoral robusticity also shows a sharp decline after the Late Paleolithic, and then
remains essentially unchanged in the sequence of the three agricultural periods. In females, no
change occurs in femoral robusticity immediately following the Late Paleolithic, but there is a
dramatic decline in the final two periods of agricultural intensification. Overall, there are
significant declines in levels of habitual activity, but with different timing between adult men
and women. This pattern fits the general finding in other regions where there is evidence for a
decline in sexual dimorphism. On the other hand, this pattern is overlain by other factors
affecting workload and activity differently for men and women that are specific to the Nile
Valley.

6.2.4 Age changes in diaphyseal structure


Via radiographic analysis, Smith and Walker (1964) demonstrated that weight-bearing long
bones (femur) undergo continuous diaphyseal expansion throughout the years of adulthood.
This pattern of bone apposition has also been observed in nonweight-bearing and weight-
bearing bones in the comparison of younger and older adults in a number of contemporary
settings (Ahlborg et al., 2003; Epker & Frost, 1966; Epker et al., 1965; Garn, 1989; Garn et al.,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
242 Activity patterns: 2. Structural adaptation

Table 6.2 Percentage change with age in femoral and tibial cross-sectional geometric
properties. Calculated by the formula: {[(40 þ years) – (20 to 39 years)]  (20 to 39 years)}
 100 (Adapted from Ruff & Hayes, 1982: Table 1)
CA MA TA Imax Imin

Bone
(section location)

Males
Tibia (20%) −1.1 13.6a 7.3a 9.4 7.8
Tibia (50%) 0.2 35.7c 12.4c 14.6a 19.9a
Femur (50%) 0.1 25.4c 6.7b 12.3b 6.8
Femur (80%) −0.5 13.1b 4.7a 7.5 4.8
Females
Tibia (20%) −14.5b 9.1a 9.1a −1.5 −3.0
Tibia (50%) −13.1b 62.4c 12.9b 10.3 4.1
Femur (50%) −6.6 66.6c 11.7c −9.6 15.6b
Femur (80%) −2.4 41.2c 13.6c 16.5b 16.0a
a
Statistically significant between age groups (Student’s t-test: P<0.05)
b
Statistically significant between age groups (Student’s t-test: P<0.01)
c
Statistically significant between age groups (Student’s t-test: P<0.001)
CA, cortical area; Imax and Imax, second moments of area/bending rigidity; MA, medullary area;
TA, total subperiosteal area.

1967, 1992; Heaney et al., 1997; Russo et al., 2006) and archaeological series (Carlson et al.,
1976; Pfeiffer, 1980; Ruff, 2010b; Ruff & Hayes, 1983b; Stirland, 1993). Some contend that
periosteal expansion represents a compensatory response to endosteal bone loss and thinning
of the cortex with advancing age (Garn et al., 1967; Ruff & Hayes, 1982; Smith & Walker,
1964; see discussion in Martin et al., 1998). Until recently, this hypothesis was difficult to test
because of the imprecision of radiographic measures of cortical bone remodeling (Ruff &
Hayes, 1982). In order to examine the issue of age changes and periosteal expansion in more
detail, Ruff and Hayes (1982, 1983b; Ruff, 2010b) analyzed section properties – areas and
second moments of area – from the late prehistoric/protohistoric Pecos Pueblo site, New
Mexico. Analysis of femoral and tibial diaphyseal sections reveals that both sexes saw
increases in medullary area and total subperiosteal area and decreases in cortical area with
advancing age (Table 6.2). Second moments of area (Imax and Imin) increase in older adults.
Thus, in support of the compensatory hypothesis, continuous periosteal expansion in older
adults appears to maintain the mechanical integrity of the long bone despite overall decline in
bone mass.
Variation in different sections along femoral and tibial diaphyses also reveals that skeletal
remodeling with age is less pronounced in the most distal and proximal sections of the
diaphyses. The greater remodeling in tibial and femoral midshaft and adjacent sections than
at the distal and proximal ends is likely due to the relatively greater mechanical loads –
especially bending (Ruff, 1992; Ruff & Hayes, 1982).
While the analysis of cross-sectional geometry and behavioral inference is richly repre-
sented in adults, it is the record of growth and development in juveniles that is important for

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 243

comprehending the acquisition of adult robusticity, the timing of its development, and the
overall record of ontogeny that informs our understanding of robusticity (Ruff, 2003).
Analysis of Medieval-period juvenile (<16 year of age) tibiae from Kulubnarti, Sudanese
Nubia, shows increases in percentage of cortical area and second moments of area (Ix, Iy) from
3 to 12 years of age (Van Gerven et al., 1985). After the age of 12 years, areas decline, while
second moments of area continue to increase dramatically. Thus, despite decline in bone area,
continued increase in second moments of area in the later juvenile years appears to preserve
mechanical integrity throughout the years of growth and development.
Similarly, analysis of various populations from North America, Europe, and the Middle East
representing a variety of behavioral, ecological, and dietary circumstances, reveals common
patterns of age changes from birth through adolescence, yet important differences reflect
significant regional variation (compare with Cowgill, 2010; Cowgill & Hager, 2007; Garofalo,
2013; Gosman et al., 2013; Larsen et al., 2013; Van Gerven et al., 1985). These comparisons
show developmental trajectories of long bone rigidity and strength. For example, in the first
year of life, Medieval Nubians and twentieth-century Portuguese show relatively low robus-
ticity, whereas Medieval Bosnians and prehistoric Inupiat show elevated robusticity (Cowgill,
2010). Sully (Arikara), Point Hope (Inuit), Çatalhöyük (Anatolian Neolithic), and Barton-
upon-Humber (Medieval-Industrial United Kingdom) express relatively high robusticity
throughout the juvenile years (Garofalo, 2013) (Figure 6.12). High levels of robusticity early
in ontogeny predict, therefore, the high levels documented for adults (Çatalhöyük;
Figure 6.13).
In order to document ontogenetic age patterns of diaphyseal remodeling in more precise
fashion, Ruff and coworkers (1994) examined areas and second moments of area of humeri in
professional tennis players aged 14–39 years. Both males and females show a pattern of
endosteal contraction and periosteal expansion resulting in large increases in bone area (CA)
and torsional rigidity (J) in the dominant playing arm. The resulting robusticity is primarily
due to greater periosteal expansion and not endosteal contraction. The degree of humeral
robusticity has a strong association with age: individuals who began playing tennis earlier
have greater robusticity than individuals who began later. The increased mechanical loading
in children and young adolescents has a more pronounced effect on the periosteal surface;
after mid-adolescence, loading appears to have a more pronounced effect on the endosteal
surface. These findings indicate an age pattern of sensitivity in the bone forming surfaces –
periosteal versus endosteal – in response to increases in mechanical stimuli. The cause of the
shift in focus of sensitivity is unknown, but may be related to changes in hormonal levels and
their different effects on the two bone surfaces (Ruff et al., 1994).
Relatively little is known about the variation in cross-sectional morphology along the
diaphyses of limb bones, especially with regard to ontogenetic changes in response to activity.
In order to draw a more precise understanding of behavioral and developmental variation,
Gosman and collaborators (2013) scanned (high-resolution X-ray CT) five sections (20%,
35%, 50%, 65%, 80%) of femoral and tibial diaphyses of juveniles ranging in age from
neonate to young adults from the Norris Farms, Illinois skeletal series. As would be predicted,
their analysis revealed that cross-sectional geometric shape (Imax/Imin) changes throughout the
life course (Figure 6.14). However, these changes are highly patterned. That is, there are peak
levels of changing morphology during two key periods of life; first, during early childhood
(attainment of walking); second, during early adolescence (hormonal and body-mass

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
244 Activity patterns: 2. Structural adaptation

(a)
10 000

9000

8000

7000
Humerus Rigidity (J)

6000

5000

4000

3000

2000

1000

0
Tigara

Tigara

Tigara

Tigara
Barton

Barton

Barton

Barton
Arikara

Arikara

Arikara

Arikara
Çatalhöyük

Çatalhöyük

Çatalhöyük

Çatalhöyük
Infant Young Child Middle Child Old Child/
Adolescent

(b)
35 000

30 000

25 000
Femur Rigidity (J)

20 000

15 000

10 000

5000

0
Tigara

Tigara

Tigara

Tigara
Barton

Barton

Barton

Barton
Arikara

Arikara

Arikara

Arikara
Çatalhöyük

Çatalhöyük

Çatalhöyük

Çatalhöyük

Infant Young Child Middle Child Old Child/


Adolescent

Figure 6.12 Developmental differences in humeral (a) and femoral (b) rigidity (J) according to
age groups and variable populations. Rigidity increases significantly from middle child to
old child/adolescent years among Neolithic Çatalhöyük, Protohistoric Arikara, and Medieval-
Industrial Barton-upon-Humber samples. Prehistoric Tigara (Point Hope) individuals do not
show this markedly increased rigidity at adolescence, indicating inter-regional variability in
long bone ontogeny. (Adapted from Garofalo, 2013; produced with permission of author.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.2 Cross-sectional geometry 245

120

Males

Females
110

100
Relative Femoral Strength

90

80

70

60
ze
ic
ic
l.
l.

h
Pa
Pa

ith

on
lit
ol

eo

Br
p.
p.

es
U
U

N
M
te
rly

La
Ea

Figure 6.13 Femoral midshaft strength (Zp) in Çatalhöyük and comparative samples. (Adapted
from Larsen et al., 2013; reproduced with permission of Cotsen Institute of Archaeology.)

Figure 6.14 Cross-sections of the femoral (above) and tibial (below) midshafts among five age
categories: Age Group 1 (0–1.9 years), Age Group 2 (2–4.9 years), Age Group 3 (5–8.9 years),
Age Group 4 (9–13.9 years), and Age Group 5 (14–17.9 years). Cross-sectional geometric
morphology (maximum diameter/minimum diameter, Imax/Imin) of the femur is generally
conserved during development while tibial midshaft becomes anterioposteriorly
strengthened. (Adapted from Gosman et al., 2013; reproduced with permission of authors
and John Wiley & Sons, Inc.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
246 Activity patterns: 2. Structural adaptation

increase). Moreover, for the tibia, the round diaphysis in earlier childhood becomes increas-
ingly anterioposteriorly strengthened. The femur, on the other hand, has more region-specific
changes, with the midshaft (50% section) being much more conserved than the distal and
proximal diaphyses. All statistically significant changes are associated with the ends of the
diaphyses, suggesting that these areas are more subject to morphological change in response
to mechanical loading during the years of growth and development. Overall, the diaphyseal
cross-sections change from relatively circular early in life to more asymmetrical in late
adolescence. This finding underscores the critical importance of understanding ontogeny as
a background to interpreting adult variation. Virtually all of the records described in this
chapter focus on adults and especially the midshaft of adult long bones. The study of Norris
Farms femoral development in particular reveals that the record of ontogeny and biomechan-
ical adaptation may be best explored and interpreted by taking a whole bone approach to
include the midshaft as well as other diaphyseal regions. Clearly, the midshaft is an important
part of the behavioral/adaptive story, but only part of the story, especially when attempting to
reconstruct key events in the lifecourse.

6.3 Histomorphometric biomechanical adaptation


Histological research on archaeological human remains focuses primarily on the documenta-
tion of systemic disturbances in interpreting remodeling patterns (e.g., nutritional depriv-
ation; see Chapter 2 and discussion in Stout, 1989; various in Crowder & Stout, 2011). Like the
overall size and morphology of skeletal elements, cortical remodeling at the microscopic level
is also influenced by the mechanical environment (Bouvier & Hylander, 1981; Stout, 1982).
Thus, the study of histological structures has considerable potential for elucidating behavioral
adaptation and activity in past human groups.
Some researchers suggest that the high levels of robusticity in Pleistocene hominin post-
crania, especially in comparison with Neanderthals and other archaic Homo sapiens with
modern humans, may be due to genetic or endocrinological differences in bone remodeling
rather than differences in mechanical loading (Abbott et al., 1996). Humerus asymmetry and
experimental evidence suggest that intrinsic factors are unlikely (see earlier; Trinkaus et al.,
1994). Microscopic analysis of Pleistocene (Neanderthal, early modern) and late Holocene
hominin (Pecos Pueblo) femora reveal no substantive differences in histological and remodel-
ing properties based on various measured and derived gross and histomorphometric param-
eters (e.g., bone area, osteon cross-sectional area, osteon cross-sectional diameter, intact
osteon population density) according to reanalysis by Streeter et al. (2010) of data originally
presented by Abbott and coworkers (1996). In virtually every parameter, archaic and modern
Homo sapiens show no difference in comparison with recent modern humans, at least as they
are compared with late prehistoric Pecos Pueblo. This finding strongly argues against genetic
or endocrinological factors in explaining high levels of skeletal robusticity in Neanderthals, in
particular, at least with respect to robusticity of the upper and lower limbs.
Comparison of histomorphometric parameters in femora of Native American (Pecos Pueblo)
and recent comparative populations (twentieth-century Euroamericans and Europeans)
reveals important patterns of variation in modern humans (Burr et al., 1990). Pecos adult
females have small Haversian canals, and males have high osteon density. Burr and coworkers
(1990) speculate that these differences reflect a more active lifestyle in Pecos Indians than in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.4 Behavioral inference from external measurements 247

other modern humans, and hence, a greater volume of bone formed per unit area. This
interpretation is in accordance with the findings based on structural and geometric analysis
of femur cross-sections in Pecos Pueblo adults (Ruff, 1991; Ruff & Hayes, 1983a, 1983b).
This reasoning is in line with the very significant positive correlations found between
osteon density and anteroposterior and mediolateral second and polar moments of inertia
in a Euroamerican cadaver sample consisting of older adults (>50 years of age); individuals
with high levels of mechanical loadings have high osteonal densities (Walker et al. 1994).
Similarly, comparisons of femoral osteon density between adult males and females from
late Christian-period Nubia (AD 550–1450) show that males have more osteons than
females, which may represent greater bone turnover from higher activity regimens in males
than in females (Mulhern, 1996). As with experimental evidence (Lanyon & Baggott, 1976),
these studies indicate that activity has a strong influence on histological variation in
cortical bone.

6.4 Behavioral inference from external measurements


Much of the record of behavioral reconstruction discussed in this chapter focuses on cross-
sectional geometric analysis obtained from the use of noninvasive technology. A key question
that has emerged is: Does the external morphology representing robusticity derived from
various diameters (e.g., anterioposterior and mediolateral diameters of long bone diaphysis)
reflect the same record of robusticity that is so well represented in cross-sectional geometric
analysis? This becomes an important question in those circumstances where the technology
for determining cross-sectional geometric properties may not be available.

6.4.1 External measurements and shapes


Overall, a number of analyses show a strong concordance between robusticity indicators
derived from external diameters and cross-sectional geometry. For example, femoral midshaft
and subtrochanteric anteroposterior and mediolateral dimensions reduce in comparisons
between prehistoric foragers and later farmers in the Georgia Bight, which Larsen (1981)
interpreted to represent decline in mechanical loading. This conclusion is supported by
analysis of second moments of area (Larsen & Ruff, 2011; Ruff et al., 1984; and compare
with Bridges, 1989b). Similarly, Great Basin foragers show high levels of robusticity based on
both approaches (Larsen, Ruff et al., 1995; Larsen et al., 2008). Statistical analysis comparing
externally derived measures of postcranial robusticity with geometrically derived measures of
postcranial robusticity in various skeletal series reveals reasonable correlations between the
two (Stock & Shaw, 2007; but see Wescott, 2006). However, the accuracy of robusticity
estimates derived from external dimensions needs to include considerations of body mass
and bone length.
Because the requirements for analysis are somewhat less stringent for external dimensions
than those in geometric analysis – the skeletal element does not have to have intact ends or
necessarily well-preserved periosteal or endosteal surfaces – a larger comparative database is
available for analysis. Comparison of hundreds (n¼524) of Archaic-period (6000–1000 BC)
and Mississippian-period (AD 1200–1600) femora from Tennessee revealed no evidence of
change in midshaft robusticity as it is expressed in mediolateral and anteroposterior

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
248 Activity patterns: 2. Structural adaptation

breadths or in the femoral robusticity index ([midshaft breadthml þ midshaft breadthap]/


femur length) (Boyd & Boyd, 1989). These findings suggest that biomechanical geometry did
not change in the transition from hunting and gathering to agriculture in this area of the
southeastern United States (compare with Bridges, 1989b). The very large published database
on external dimensions allows for the inclusion of more individuals and thus a more
complete assessment of variability within human populations than is normally possible
from cross-sectional geometric analysis. Variation showing increases and decreases in
robusticity in the Old World similarly reveals the different responses to dietary shifts,
especially with regard to the shift from foraging to farming or agropastoralism (compare
with Jacobs, 1993; Smith, Bloom et al., 1984). The role of labor and skeletal robusticity
based on external dimensions is especially well illustrated in historical contexts involving
known labor and activity regimens. For example, members of society living in poor houses
during the nineteenth century in the American Northeast express quite elevated levels of
skeletal robusticity, a finding consistent with historical records documenting heavy labor
demands (Phillips, 2003).
External bone dimensions and comparisons of left and right humeri can be used to infer
differential use of the arms, albeit with less precision than available from geometric analysis
(Borgognini Tarli & Repetto, 1986; Stirland, 1993). Comparison of paired male humeri from
two Medieval British skeletal series, Norwich and Henry VIII’s flagship, the Mary Rose,
provides a perspective on upper limb use (Stirland, 1993). Norwich males possess marked
humeral asymmetry with clear right dominance, whereas Mary Rose males exhibit very little
evidence of asymmetry. Mary Rose males exhibit a pronounced hypertrophy of the greater
tubercle on the left humerus. These findings suggest that, in contrast to Norwich males, the
Mary Rose males subjected their left and right arms to relatively equal mechanical loads
(Stirland, 1993). This finding is compatible with the historical records indicating that many of
the deceased from the Mary Rose were professional archers, an activity requiring great
strength in both arms.
That higher-status individuals may be subject to less of a workload than lower-status
individuals has been tested by comparison with external diaphyseal skeletal robusticity in
the Oleniostrov Russian Mesolithic series (Jacobs, 1995). In this setting, adult males from the
artifactually “richest” graves show the least amount of humeral and femoral robusticity. Thus,
wealthier, high-status individuals are less robust than poorer individuals. This finding sug-
gests that accumulation of wealth in this society was not achieved by having great musculo-
skeletal strength, but that it did confer socioeconomic benefits such as a reduction in the
amount of the workload.
External diaphyseal shape differences between human groups have also been documented.
Various researchers report a temporal increase in circularity of the femur (subtrochanteric
and midshaft regions) or tibia (midshaft), or both skeletal elements, with the transition to
sedentary lifeways, especially in settings involving the adoption of agriculture in the
American Southwest (Bennett, 1973), Southeast (Hoyme & Bass, 1962; Larsen, 1982,
1984), and Midwest (Perzigian et al., 1984). Increasing circularity of lower limb bones is
apparently a worldwide trend since the late Pleistocene (Anderson, 1967; Brothwell, 1981;
Buxton, 1938; Smith & Jones, 1910; K. Kennedy, 1989; Kimura & Takahashi, 1982; Larsen,
1982; Lovejoy et al., 1976; Ruff et al., 1984; Townsley, 1946; Wanner et al., 2007). With
regard to the femoral and tibial midshaft diaphyses, the respective pilasteric and platycnemic

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.4 Behavioral inference from external measurements 249

indices (both computed by the formula: [breadthml/breadthap]  100) tend to be lower in


more mechanically stressed populations than in less mechanically stressed populations.
More stressed populations possess greater mediolateral flattening than less stressed popula-
tions, which may indicate relatively greater anteroposterior bending forces in the upper and
lower legs (Lovejoy et al., 1976; Ruff & Hayes, 1983a; Ruff et al., 1984). For the pilasteric
index, Ruff (1987, 1992) showed a decline in degree of sexual dimorphism in comparing
prehistoric hunter-gatherers, prehistoric agriculturalists, and industrial Western populations,
which closely follows the patterns observed for the geometric Ix/Iy ratio in these same
populations (see earlier). Ovoid or flattened cross-sections of long bones were previously
interpreted to reflect responses to suboptimal nutrition (Angel, 1979; Buxton, 1938;
Hoyme & Bass, 1962). Analysis of bone areas and second moments of area does not support
this interpretation.
Historically, a number of researchers did invoke functional arguments to explain the
flattening of long bone diaphyses, such as with regard to the effects of specific muscles or
muscle groups on diaphyseal morphology (Angel, 1971a; Chesterman, 1983; Fowke, 1902;
Matthews, 1893). For example, Angel (1971a) related anteroposterior flattening of the
femoral proximal subtrochanteric diaphysis in pre-Classical Lerna, Greece to greater stresses
exerted by “gluteal and other hip-balancing” muscles. Similarly, he indicated that medio-
lateral flattening of the femoral midshaft was due to actions of the quadriceps muscles.
Although these muscle groups contribute to bending and torsional loading, cross-sectional
geometric analysis indicates that shaft shape is primarily influenced by bending and
torsional forces acting on the entire diaphysis and not direct actions of specific muscles or
muscle groups.
Clearly, externally defined long bone diaphyseal size and shape based on linear dimensions
provide insight into biomechanical function, but it is important to emphasize that cross-
sectional geometry is a far more precise measure, especially because it gives details on the
distribution of skeletal tissue in a section.
This perspective also demonstrates the limitations of interpreting function from measure-
ment of cortical bone thickness alone or other indictors of bone mass. For example, relative
thickness of cortical bone quantified by various versions of the percentage of cortical area
index has been widely used as an indicator of nutritional health; for example, low values of
the percentage of cortical area in relation to some standard are interpreted to represent a
deficiency in nutrition status (Brown, 1988; Cook, 1984; Storey, 1992a; and see review in
Pfeiffer & Lazenby, 1994), and imply negative effects on bone structure. Some researchers
relied on cortical thickness or mass as an indicator of functional demand alone (Hatch et al.,
1983; Smith, Bloom et al., 1984). The percentage of cortical area could reflect several things,
including a relatively small medullary space, or an expansion of the periosteum, or even a
combination of both medullary contraction and periosteal expansion (see discussions in
Ruff & Larsen, 1990; Ruff et al., 1994). Relatively thick cortical bone in a section could even
be associated with comparatively reduced bone rigidity (Figure 6.15). Ruff and coworkers
(1984) found that late prehistoric populations from the Georgia coast have among some of the
highest values of percentage of cortical area in modern humans, yet the cortical tissue is
tightly constricted about the central axis, thus resulting in relatively low values of bone
rigidity. For this reason, mechanical analysis and behavioral inference should be determined
from both bone areas and second moments of area.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
250 Activity patterns: 2. Structural adaptation

% Cortical Thickness: - 33%

% Cortical Area: - 33%

Cortical Area: +12%

Second Moments of Area: +100%


Figure 6.15 Effects of subperiosteal and medullary expansion on cross-sectional geometry.
The section on the right has a reduced percentage cortical thickness and percentage
cortical area compared to the section on the left. Cortical area (reflecting axial strength)
increases slightly and second moments of area (bending strength) increase dramatically.
(From Ruff, 1992; reproduced with permission of author and John Wiley & Sons, Inc.)

6.4.2 Populations on the move: mobility and skeletal robusticity


Diaphyseal form in cross-section is a sensitive record of habitual direction of mechanical
loading. A growing body of experimental evidence shows that long bone diaphyses respond in
predictable ways, specifically increasing amount of cortical bone in the primary plane of
deformation in bending (Lanyon, 1992; Macdonald et al., 2009; Shaw & Stock, 2009b). For
example, the comparison of boys (ages 9–11 years) engaging in exercise involving anterio-
posterior loading of the tibia and a control group revealed greater increases in tibia diaphyseal
cortical area and cortical thickness in the anterioposterior plane than in the mediolateral plane
in the exercisers, along with greater increases in Imax, oriented more anterioposteriorly, than
in Imin (Macdonald et al., 2009). Similarly, the comparison of tibia diaphyses of male univer-
sity cross-country runners and field hockey players showed greater values of second moments
of area (J, Imax, Imin) and cortical area than controls not involved in sports or other strenuous
activities (Shaw & Stock, 2009b). The runners had higher Imax values in the anterioposterior
plane than the hockey players and the controls, and more elongated cross-sections with
greater diameters. These differences likely reflect the strong emphasis on unidirectional
locomotor action in the runners in comparison with hockey players, who had rounder bones.
Overall, these studies support the notion that bone cross-sectional shape as well as size reflects
habitual activity.
Analysis of cross-sectional shape in archaeological and other skeletal series based on the ratio
of Ix/Iy in the femoral midshaft is consistent with this experimental evidence for athletes.
Specifically, an Ix/Iy ratio close to 1.0 reflects a nearly circular shape, and a ratio deviating
from 1.0 represents an ovoid shape (Figure 6.16). This ratio for the femur midshaft assesses the
distribution of bone in the mediolateral and anteroposterior planes; an ovoid-shaped midshaft

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.4 Behavioral inference from external measurements 251

Ix

Ix

Iy Iy

Ix/Iy = 1.0

Ix/Iy = 1.5

Figure 6.16 Schematic representation of femoral rigidity based on midshaft dimensions.


The left cross-section represents a femur where less anterioposterior mechanical loading
has occurred (Ix/Iy¼1). This ratio is more consistent with a sedentary lifestyle. The right
cross-section shows a femur where significant anterioposterior loading has occurred. This
cross-sectional mobility index (Ix/Iy >1) is consistent with more mobile populations.

in the anteroposterior plane (i.e., ratio is >1.0) represents relatively greater bone rigidity and
functional demand in the anteroposterior direction than in the mediolateral direction. Ruff
(1987) has shown a temporal decline in the ratio in recent human groups, which he interpreted
to reflect a general reduction in the amount of anteroposterior bending forces as populations
have become increasingly sedentary (compare with Lovejoy et al., 1976). Therefore, the Ix/Iy
ratio represents a mobility index, herein referred to as Ruff’s mobility index.
Analysis of the generally high mobility index in hunter-gatherers from the American Great
Basin, especially from Stillwater and the Great Salt Lake regions, reveals the dominance of a
highly mobile lifestyle, especially in males (Ruff, 1999). Their values are among the highest in
comparison with other North American populations, and like data presented for second
moments of area (J), the mobility index values fit into a continuum from generally high in
hunter-gatherers to relatively lower in agriculturalists. Based on this index, Great Basin
females appear to be less mobile than males, which is true for all other North American series
(Ruff, 2010b). Comparisons of females and males in the Ix/Iy ratio show that Great Basin
foragers are highly sexually dimorphic (Figure 6.17), which is a paramount characteristic of
hunter-gatherers generally, and is related to sexual division of labor and a strong male–
female dichotomy in activity patterns (Larsen, Ruff et al., 1995; Ruff, 1987, 2010a; Ruff &
Larsen, 2014; Stock et al. 2011).
In the Georgia Bight, early contact-period femora show a general reduction in the femoral
midshaft Ix/I y ratio relative to earlier prehistoric populations. Historic sources indicate that
populations during the historic period became generally less mobile as they were either forced
or coerced to live in and around mission centers (e.g., Santa Catalina de Guale, Georgia). This
skeletal indicator of mobility is, therefore, in accordance with other sources describing
substantial population sedentism during the historic period. These structural modifications
suggest that mission Indians in Spanish Florida had a more physically demanding lifestyle,
but within the relative confines of the mission setting (Larsen & Ruff, 1994).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
252 Activity patterns: 2. Structural adaptation

30

Alabama
Georgia-Florida

25 Other North America

20
Sexual Dimorphism in Ix/Iy (%)

15

10

Hunter-Gatherers Agriculturalists
0
Pecos
Malheur
Stillwater

Delaware

Guale (HG)

Guale (AG)
Alabama (HG)

Alabama (AG)
E. Gr. Salt Lake

New Mexico (HG)

New Mexico (AG)


Gr. Plains Lt. Coal.
Gr. Plains Er. Coal.

Late Mission Guale


Gr. Plains Pre-Coal.

Early Mission Guale

Early Mission Timucua

Figure 6.17 Mobility indices for multiple preagricultural and agricultural populations
indicate an appreciable percentage of sexual dimorphism in comparison to hunter-
gatherers and horticulturalists in North America. However, intensive labor exploitation
among later Spanish mission Indians in Guale indicate great differences between males and
females. (From Larsen & Ruff, 2011; © 2011 John Wiley & Sons, Ltd.)

Similarly, temporal comparisons of Early Classic (AD 250–550) and Late Classic
(AD 550–750) populations from Xcambó on the northern coast of the Yucatán peninsula,
Mexico revealed a significant decline in the Ix/Iy ratio in males (but not females), indicating
considerable decline in mobility (Maggiano et al., 2008; and see Wanner et al., 2007). In this
setting, there was no clear change in diet (Maggiano et al., 2008). Rather, the change
in activity – also accompanied by a general decline in robusticity – appears to be related to
improvements in general living standards. In particular, evidence for improvements in
material wealth (e.g., architecture, grave inclusions) suggests that increasing economic

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.4 Behavioral inference from external measurements 253

complexity associated with the administrative function of the community resulted in reduced
mobility and workload in general, especially in males. These findings are consistent with other
circumstances showing a negative association between wealth and workload in stratified
societies (Robb et al., 2001).
An important European record of diaphyseal shape ratios using cross-sectional shape
reveals similar trends in activity. In a series of femora from multiple localities representing
the Late Upper Paleolithic, Mesolithic, Neolithic, Eneolithic, and Iron Age (c. 19 000–4000
yBP) in northwestern Italy, analysis of the Imax/Imin ratio (the ratio of plane of maximum
bending strength to plane of minimum bending strength) shows a continuous decline in
females, reflecting an increasingly round shape in cross-section of the femoral diaphysis
(Marchi, 2008; Marchi et al., 2011; Sparacello et al., 2011). This expected finding contrasts
with the record of the Imax/Imin ratio for males. That is, following the Mesolithic, Neolithic
males show a dramatic rise in the ratio, approaching the mean for males of the Late Upper
Paleolithic. This suggests that the Neolithic males, but not females, were likely as mobile as
Late Upper Paleolithic males, certainly more so than the Mesolithic and all post-Neolithic
populations in the region (Marchi et al., 2011; and compare with Sládek et al., 2006a, 2006b
for central Europe). Overall, therefore, these findings point to an exception to the general
finding of declining mobility from the late Pleistocene through recent times. What is it that
Neolithic males from northwest Italy are doing that requires elevated mobility? Marchi and
coworkers (2011) suggest that the spike in mobility in Neolithic males reflects the emphasis on
pastoral activities, especially animal husbandry, and not agricultural/farming activities. For
example, Robb (1994) makes the argument that these Neolithic males were involved in
herding over rough and uneven terrain. Indeed, the biomechanical evidence strongly supports
this explanation. This pattern of elevated mobility contrasts with patterns of the probability
ratio for Neolithic Çatalhöyük (Larsen et al., 2013) (Figure 6.18). That is, while the ratio
documenting the anterioposterior to mediolateral bending strength is relatively high (>1.0), it
is considerably lower than samples documenting the European Upper Paleolithic, Mesolithic,
Neolithic, and Bronze Age. Thus, while physically active, the relatively lower mobility index
value for Çatalhöyük is more consistent with a population that is relatively sedentary. On the
other hand, within the 1400-year period of occupation, there is a trend showing an increase in
the index, suggesting that the population became increasingly mobile, perhaps in response to
a decrease in food and other resources immediately adjacent to the community, in combin-
ation with increased aridity prior to the abandonment of the community, as well as a range of
other factors (Larsen et al., 2013) (Figure 6.19).

6.4.3 Femoral neck–shaft angle


The femoral neck–shaft angle is a measurement of the relative degree of more medial versus
more proximal orientation of the femur neck to the diaphysis. In modern adults, the range is
variable, from 110 to 150 (Anderson & Trinkaus, 1998; Trinkaus, 1993). Active juveniles
show a greater decrease in the angle during the years of growth and development than
inactive juveniles (Houston & Zaleski, 1967). The orientation of the femoral neck relative to
the diaphysis appears to be responsive to the combined forces of body weight, muscle forces,
and activity generally. Additionally, the smaller angle provides greater hip joint mechanical
stability under increased mechanical loading (Trinkaus, 1993).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
254 Activity patterns: 2. Structural adaptation

1.6

Males

Femoral A-P/M-L Bending Strength Females


1.4

1.2

1.0

0.8
.

l.

ic

ze
l
Pa

Pa

hi
ith

on
lit
ol
p.

p.

eo

Br
es
U

N
M
rly

te
La
Ea

Figure 6.18 Mobility indices (femoral A-P/M-L bending strength) of Neolithic Çatalhöyük
samples compared with Early Upper Paleolithic, Late Upper Paleolithic, Mesolithic,
Neolithic, and Bronze Age European populations. Çatalhöyük males and females have
mobility indices close to 1.0, suggesting a more sedentary lifestyle relative to these
European sites. (Reproduced with permission of Cotsen Institute of Archaeology.)

1.3

Males
1.2
Femoral A-P/M-L Bending Strength

Females

1.1

1.0

0.9

0.8

0.7
Early Middle Late
Figure 6.19 Mobility indices (femoral A-P/M-L bending strength) for Neolithic Çatalhöyük
indicate an increase in mobility from Early to Late periods in both males and females.
(Reproduced with permission of Cotsen Institute of Archaeology.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
6.5 Summary and conclusions 255

Comparisons of a range of human groups – foragers, agriculturalists, and urban dwellers –


support the mechanical hypothesis for variation in the femoral neck–shaft angle (Trinkaus,
1993). In particular, foragers have the lowest femoral neck–shaft angles (mean¼125.7),
urban dwellers have the highest angles (mean¼132.3), and agriculturalists lie between the
foragers and urban samples (mean¼128.2) (means calculated from Trinkaus, 1993: Table 4).
This pattern of increasing neck–shaft angles from mobile foragers to sedentary urban groups
closely parallels general trends of decreasing robusticity based on whole bone measurements
and diaphyseal structure (Ruff et al., 1993), albeit with a wide range of overlapping variability
between groups (Trinkaus, 1993).

6.5 Summary and conclusions


Limb bone diaphyses are highly responsive to the mechanical environment. Structural and
histomorphometric analysis from a variety of settings underscores the extraordinary develop-
mental plasticity of bone tissue throughout the lifespan, thus providing an important record of
magnitude and types of mechanical loading. Differences between human populations provide
insight into behavioral patterns in the past. In contrast to diaphyses, articular size and
morphology appear to be resistant to mechanical demand, which reflects the strong genetic
canalization regarding the morphology of the articular joints.
Because humans normally do not use their upper limbs in ambulatory activities, the influ-
ence of body weight is minimal in the overall determination of size and morphology of arm
skeletal elements (e.g., humerus). The study of upper limb bilateral asymmetries in various
human populations allows inferences to be made about loading levels and patterns in relation
to different lifestyles and mechanical functions.
Measurement of areas and second moments of area in a range of human populations,
ancient and modern, reveals a general trend for decline in robusticity and mobility. Although
this trend is especially pronounced in the transition from archaic to modern Homo sapiens in
the late Pleistocene (Holt, 2003; Holt & Formicola, 2008; Ruff et al., 1993; Shackelford, 2007;
Sparacello et al., 2011), it has continued throughout the Holocene in the foraging-to-farming
transition. The reduction during the Holocene is likely tied to increasing sedentism associated
with plant domestication and increasing technological efficiency (Larsen, 1995). Some
modern human populations are quite robust (e.g., early modern Europeans, Neolithic Italians,
Great Basin Native Americans), which may be linked with living in marginal environments
and the great physical effort in the quest for food in these types of settings.
External shaft dimensions, measures of bone mass or volume (e.g., percentage of cortical
area), and femoral neck–shaft angles provide some insight into activity patterns, but precision
in behavioral interpretation is dependent upon analysis of skeletal tissue distribution via
cross-sectional geometry. External dimensions are somewhat limited in estimation of robus-
ticity, but with appropriate size control, they provide a rough substitute to cross-sectional
geometry. The relative degree of flattening of long bone diaphyses, especially in the proximal
and midshaft femur and midshaft tibia, is related to type and level of mechanical loading and
not to nutritional factors or the actions of specific muscle groups.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 14:02:04, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.008
Masticatory and nonmasticatory
functions: craniofacial adaptation to
7 mechanical loading

7.1 Introduction
The influence of environment and behavior on skull morphology was first discussed centuries
ago. In the fifth century BC, Herodotus remarked on apparent differences in cranial robusticity
between Persians and Egyptians: “The skulls of Persians are so weak that if you so much as
throw a pebble at one of them, you will pierce it; but the Egyptian skulls are so strong that a
blow with a large stone will hardly break them.” He interpreted the strength of Egyptian skulls
to the lifelong exposure of their heads to the sun. Since the time of Herodotus, the influence of
environment and behavior, no matter how specious the interpretation, has been only minim-
ally considered in discussions of cranial morphology in archaeological remains. Beginning in
the eighteenth century, osteologists relied on craniofacial variation for determining popula-
tion history and classification, with little attention given to the biological significance of
observed patterns, especially from a population perspective (Armelagos et al., 1982; Carlson &
Van Gerven, 1979; Cook, 2006; Larsen & Walker, 2010; Powell, 2005). As with the study of
long bone morphology discussed in Chapter 6, there has been a gradual reorientation from
typological/historical to population variation analyses. This emphasis focuses on processes
and functions that influence cranial morphology, revealing the adaptive and behavioral
significance of variation from a population perspective.
This chapter’s central discussion is built on the premise that craniofacial morphology is a
biological ensemble, influenced by processes involving the central role of the masticatory
complex – composed of the jaws, teeth, and associated muscle – bone interactions – during
growth and development and throughout adulthood (Daegling, 2010; Lieberman, 2011). Both
intrinsic and extrinsic factors influence craniofacial growth, just as with any other skeletal
tissue in the body. Throughout life, the cranium grows and adapts in an epigenetic context
where environmental and functional factors operate on an underlying genetic infrastructure
(Carlson, 2005).

7.2 Cranial form and functional adaptation


7.2.1 Determinants of form
Cranial form in the growing child and the maturing adult is determined by a complex
interaction of intrinsic (hormones, genetic) and extrinsic (environmental, especially mechan-
ical) factors. The overall form is principally a product of natural selection (compare with
Herring, 1993; Maynard Smith & Savage, 1959). Animal and human heritability demonstrates
the strong genetic component of craniofacial form (Atchley, 1993; Carson, 2006; Devor, 1987;
Johannsdottir et al., 2005; King et al., 1993; Savoye et al., 1998), but experimental and

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.2 Cranial form and functional adaptation 257

observational studies on animals and humans (Daegling, 2010; Herring, 1993; Kiliaridis,
1995; Lieberman, 2011; Raadsheer et al., 1999; Ravosa et al., 2008) show the considerable
influence of environment, especially in relation to the cumulative effects of mastication and
mechanical loading of the face and jaws during the years of growth and development and
adulthood. Cranial form cannot be regarded as a single unit with similar genetic and environ-
mental influences. That is, regions of the cranium that are subjected to masticatory loading
(e.g., attachment sites for masticatory muscles in the face and jaws) show more variation than
other regions of the cranium (Lieberman, 1995; von Cramon-Taubadel, 2009b).
Experimental studies involving extirpations of masticatory muscles in laboratory animals
show associated craniofacial skeletal modifications, especially with regard to size and robus-
ticity (Avis, 1959, 1961; Horowitz & Shapiro, 1951, 1955; Moore, 1967, 1973; Pratt, 1943;
Schumacher et al., 1979; Washburn, 1947a, 1947b). Craniofacial skeletons of animals fed soft
foods tend to be smaller and less robust than animals fed hard foods (Beecher & Corruccini,
1981, 1983; Bouvier & Hylander, 1981, 1982, 1984; Bouvier & Zimny, 1987; Corruccini &
Beecher, 1982, 1984; Hinton, 1990; Lieberman, Krovitz et al., 2004; Moore, 1965; Tuominen
et al., 1993; Watt & Williams, 1951; Whitley et al., 1966; and see reviews by Herring, 1993;
Kiliaridis, 1995; Lieberman, 2011). The profound effects of alteration of masticatory loading
are also demonstrated in the experimental transpositions of masseter and temporalis muscles
in laboratory monkeys (Hohl, 1983). The anterior relocation of these muscles leads to a
number of changes, including superior facial tilting. These extirpation and translocation
studies show that alterations in mechanical loading produce shifts in masticatory behavior
that result in distinctive craniofacial morphological changes. Collectively, this body of work
underscores the centrality of adaptive plasticity, or the organism’s ability to respond to
environmental circumstances during their ontogeny (Gotthard & Nylin, 1995; Ravosa et al.,
2008; and see Chapter 6).

7.2.2 Temporal trends in human populations


Contrary to the assertion that human head form is stable and highly heritable (Dixon, 1923;
Neumann, 1952; and see Gould, 1996; Marks, 1995), diachronic population studies reveal a
high degree of plasticity. Franz Boas was among the first to demonstrate clear evidence of
secular change in head shape. Based on a ratio of head length to breadth (cephalic or cranial
index), American-born immigrants are appreciably different in cranial form from their
European-born parents (Boas, 1912, 1916; also Hrdlička, cited in Boas, 1916:716). Based on
this observation of plasticity, Boas was strongly opposed to the use of cranial form for tracing
population history and linking past with living groups. Rather, he argued that “the anatomical
forms of the present population and of ancient skeletons do not allow us to draw inferences
regarding nationality of the ancient inhabitants” (Boas, 1902:445). The high degree of
developmental plasticity in craniofacial form among twentieth-century populations was
confirmed further by Shapiro and Hulse’s comparisons of Japanese born in Hawaii and
Japanese immigrants to Hawaii (Shapiro, 1939). The differences between the two were
pronounced; the longer the time the immigrant population was living in Hawaii, the greater
the differences compared with the ancestral population still living in Japan. Similarly, change
in head shape in other recent populations appears to reflect local environmental and cultural
circumstances and the plasticity of cranial form, including increased vault height and change

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
258 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

White males
144
White females
Black males
142
Black females
Basion-Bregma Height 140
138
136
134
132
130
128
126
124
122

1840 1860 1880 1900 1920 1940 1960 1980


Decade of Birth
Figure 7.1 Diachronic changes in basion-bregma height among White males (circles), White
females (triangles), Black males (squares), and Black females (diamonds) from 1840 to 1980.
(From Jantz & Meadows Jantz, 2000; reproduced with permission of authors and John
Wiley & Sons, Inc.)

in width (Buretic-Tomljanovic et al., 2006; Cameron et al., 1990; Jantz, 2001; Jantz & Logan,
2010; Jantz & Meadows Jantz, 2000; Weisensee & Jantz, 2011) (Figure 7.1). Simultaneous
increases in head length and long bone length documented in North Americans of European
and African ancestry suggest that these are responses to similar forces affecting growth (Jantz
& Meadows Jantz, 2000).
An independent approach to understanding plasticity of cranial shape involved efforts to
document the influence of culture and behavior in past populations. In sharp contrast to
ancient British populations, Keith observed that in modern Britons “many persons have small,
contracted palates. . .Their noses are narrow; so are their faces” (1950:402). Late Celtic faces
became smaller, reflecting in part the “change in dietetic which has occurred since the early
years of the Christian era, cooked food and soft cereals replacing tough meats and imperfectly
ground corns” (Keith, 1916:198). Keith (1916) noted concomitant changes in the occlusal
surfaces of teeth, especially with regard to a reduction in tooth wear in later populations.
Keith’s observations regarding the influence of dietary consistency on facial robusticity were
prescient of a large body of work showing the effects of cooking and reduction of food and
links with reduction of growth of the masticatory apparatus both in Britain (Frake & Goose,
1977; Goose, 1962, 1981; Goose & Parry, 1974; Lavelle, 1972; Moore et al., 1968) and
elsewhere.
Following Keith’s efforts to relate temporal trends in craniofacial morphology to shifts in
masticatory behavior and diet, various researchers have documented other trends in archaeo-
logical populations, both regionally and globally. Weidenreich (1945) recognized the
inappropriateness of using cranial shape for identifying racial groups, instead observing a

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.2 Cranial form and functional adaptation 259

trend in human evolution – and especially during the Holocene – for increasingly shorter
crania, a process he called brachycephalization. Most workers argued that long-headed
(“dolichocephalic”) populations had been replaced by alien short-headed (“brachycephalic”)
populations, thus explaining the trend (Retzius, 1900; and see review in Weidenreich, 1945).
Contrary to this consensus, Weidenreich (1945) argued against invasion and replacement
models by showing the widespread trend of cranial vault shortening, taking place in earlier
populations throughout Europe (Sokal & Uytterschaut, 1987; Sokal et al., 1987), the Middle
East, and South and Central Asia (Vladescu, 1992).
As with other regions of the world, studies of native New World groups in the first half of
the twentieth century emphasized diffusionistic interpretations of cranial shape variation,
especially arguing that earlier long-headed “dolichocephals” were replaced by later “brachy-
cephals” in native New World groups (Dixon, 1923; Hooton, 1930, 1933; Hrdlička, 1922a;
Hulse, 1941; Newman & Snow, 1942).
Coinciding with and following the publication of Weidenreich’s (1945) classic article, other
researchers recorded temporal changes in cranial form, especially involving reductions in
robusticity and/or increasing vault roundness and brachycephalization, in the comparison of
earlier and later prehistoric North American Indians (Anderson, 1967; Boyd, 1988; El-Najjar,
1981; Guagliardo, 1982b; Hoyme & Bass, 1962; Ivanhoe, 1995; Newman, 1962; Newman &
Snow, 1942; Steele & Powell, 1992; Webb & Snow, 1945) and elsewhere around the globe
(Abdushelishvili, 1984; Henke, 1984; Kaifu, 1997; Mushrif-Tripathy & Walimbe, 2006; Naka-
hashi, 1993; Newman, 1951; Okazaki, 2004; Rightmire, 1984; Rösing & Schwidetzky, 1984;
Sagne, 1976; Sardi et al., 2004; Smith, Bar-Yosef et al., 1984; Smith & Horwitz, 2007; Suzuki,
1969; Walimbe & Gambhir, 1994; Walimbe & Kulkarni, 1993; Walimbe & Tavares, 2002;
Wu & Zhang, 1985; Wu et al., 2012; and references cited later). In some settings, cranial vault
shortening can also be attributed to artificial deformation in some cultures, a practice largely
limited to late prehistoric populations. For much of this record, there is a general pattern of
gracilization of the face and jaws in comparing hunter-gatherers and early farmers in various
settings (Bulbeck & Lauer, 2006; Larsen, 1982, 1995; Sardi & Béguelin, 2011; Sardi et al.,
2004; von Cramon-Taubadel, 2011).

The Nile Valley: invasion and replacement or subsistence change and craniofacial adaptation?
Beginning in the nineteenth century, various workers speculated on the origins of human
groups occupying the Nile Valley (Morant, 1925; Morton, 1844; Smith, 1910). Following
Morton’s (1844) highly influential study of archaeological crania from Egypt and Nubia, the
prevailing notion was that two biologically distinct groups occupied the Nile Valley in
temporal succession. In Lower Nubia, Morant (1925) identified an earlier “Upper Nile type,”
with predominantly “Negroid” features, and a later “Lower Nile type,” which lacked “Negroid”
features. The changes were viewed in a diffusionistic paradigm: simply, the disappearance of
“Negroid” features resulted from an invasion and subsequent replacement by alien “Caucas-
oid” (Egyptian) peoples from the north (Calcagno, 1986a, 1986b; Carlson, 1976; Carlson &
Van Gerven, 1977, 1979; Van Gerven et al., 1973, 1977).
Recent analyses of crania and dentitions from Lower Nubia indicate that the evidence for
the diffusionist model of biological change is less than compelling. Independent analyses of
skeletal and dental discrete and metric variables and other lines of evidence suggest that the
earlier and later Nubian populations largely represent a biological continuum (Batrawi, 1946;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
260 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

Berry & Berry, 1972; Calcagno, 1986a, 1986b; Franceschi et al., 1994; Greene, 1972; Mukher-
jee et al., 1955; Nielsen, 1970; Van Gerven et al., 1977). Certainly, there was considerable
contact and likely gene flow between Nubia and other settings in northeast Africa, as is
clearly illustrated in the long history of contact between Nubia and Egypt (see Chapter 4).
However, there is no evidence of population replacement. Therefore, the differences in cranial
morphology between earlier and later populations observed by Smith and Jones (1910),
Morant (1925), and others are best understood in relation to factors not involving population
replacement.
In order to understand these factors better, especially those related to dietary and
technological change, Carlson and Van Gerven and their coworkers (Armelagos et al.,
1984; Carlson, 1976; Carlson & Van Gerven, 1977, 1979; Hinton & Carlson, 1979; Van
Gerven et al., 1973, 1977) compared craniofacial morphology in a Nubian-based temporal
sequence, including foragers from the Mesolithic (c. 12 000 yBP), initial agriculturalists
from the combined A- and C-groups (3400–1200 BC), and intensive agriculturalists from
the combined Meroitic, X-group, and Christian periods (AD 0–1500). These comparisons
reveal that Nubian foragers and incipient farmers have flat and elongated vaults with
well-developed, protruding supraorbital tori and occipitals. In contrast, later intensive
agriculturalists have rounded vaults with small and more posteriorly positioned faces and
masticatory muscle (temporalis and masseter) attachment sites and reduced temporoman-
dibular joint size (Figure 7.2).
Carlson and coworkers posit a masticatory-functional hypothesis for explaining cranio-
facial changes in Nubia (Figure 7.3). They argue that the primary factor influencing Nubian
craniofacial anatomy was the change in subsistence economy, from foraging to food produc-
tion and the shift to consumption of softer foods. These changes resulted in a reduction in
loading of the masticatory muscles. Alteration in masticatory function led to alteration in
craniofacial growth in two ways: (1) decreased stimulation of bone growth leading to a
reduction in facial robusticity, and (2) progressive alteration of the overall growth of the face
and vault, resulting in a smaller and more inferoposteriorly oriented face relative to the
cranial vault.

Global patterns of biocultural adaptive plasticity


The masticatory-functional interpretation of diachronic change in craniofacial morphology in
Nubia offers an important means of interpreting morphological changes elsewhere, especially
where food production and agriculture have supplanted hunting and gathering as a primary
mode of subsistence. Neolithic mandibles from Lepenski Vir, Vlasac, and Vinca in the Balkans
region of central Europe show a reduction in size in comparison with earlier Mesolithic
mandibles (y’Edynak and Fleisch, 1983). In this region, the change in size of the mandible
coincides with the shift from foraging and fishing in the earlier period to farming of several
grains (e.g., eincorn and emmer wheat). Unlike the foraging adaptation, the latter dietary focus
also involved extensive cooking of food in ceramic vessels. The mastication of generally softer
foods, therefore, resulted in substantial decreases in mechanical loading of the craniofacial
skeleton. This conclusion is supported by comparisons of gross morphology and cross-sectional
geometry of mandibular corpi of juvenile and adult foragers having heavy masticatory
loading (Point Hope Inuit; AD 1200–1700) with juvenile and adult farmers consuming

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.2 Cranial form and functional adaptation 261

Figure 7.2 Summary of craniofacial changes in Nubia, comparing Mesolithic foragers


with Meroitic-Christian farmers (dashed line). Note that farmers have relatively
greater posterior placement of areas of muscle attachment, facial reduction, vault
length reduction, vault height increase, and more globular shape than foragers.
(From Larsen, 2014; adapted from Carlson & Van Gerven, 1977; reproduced with
permission of authors, John Wiley & Sons, Inc., and W.W. Norton and Company.)
(A black and white version of this figure will appear in some formats. For the color
version, please refer to the plate section.)

softer foods (Sully Arikara; AD 1650–1750) (Holmes & Ruff, 2011). In this regard, the Point
Hope corpi are considerably more buttressed mediolaterally than Sully site corpi. Moreover,
second moments of area (Ix) of mandibular corpi are significantly greater in the Point Hope
series than in the Sully series overall (see Chapter 6 for discussion of cross-sectional geometry).
Interestingly, the differences emerge in young juveniles (aged 1–6 years), and become statistic-
ally significant in later childhood (6–12 years old), adolescence (12–18 years old), and adult-
hood (post-18 years of age). Similarly, comparisons of mandibular robusticity in prehistoric
Japan reveal differences in comparison with modern Japanese (Fukase & Suwa, 2008; Kaifu,
1997; Okazaki, 2004), patterns that emerge early in childhood. These examples show that
differences in robusticity that develop through ontogeny meet expectations of reduced masti-
catory loading. The evidence for either development of robusticity or reduced robusticity during
the growth years also lends considerable support for adaptive plasticity in explaining these
patterns of variation.
Temporal trends in skeletal series from the Eastern Woodlands of North America are
consistent with the masticatory-functional paradigm. Crania of late prehistoric (Mississippian
period) farmers from Tennessee show a general decrease in robusticity and a reduction in size
and more posterior orientation of the masticatory muscles in comparison with crania of early

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
262 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

Foraging to farming

Reduced toughness of food

Altered chewing

Reduced masticatory
muscle activity

Altered craniofacial growth

Reduced Reduced size and


facial growth configuration of
masticatory muscles

Reduced Decreased stimulation


face and jaws of bone growth in
face and jaws

Reduced robusticity of
face and jaws

Response of cranial vault and


reduced amount of bone

Shorter, wider, more globular


vault with less-projecting
face and less room for teeth
(increased malocclusion)

Figure 7.3 Masticatory-functional model of craniofacial change in Nubia. (From Larsen,


2014. © 2014, 2011, 2008 by W. W. Norton & Company, Inc. Used with permission of W. W.
Norton & Company, Inc.) (A black and white version of this figure will appear in some
formats. For the color version, please refer to the plate section.)

prehistoric (Archaic period) foragers (Boyd, 1988). In this setting, size reduction is especially
pronounced in the mandible and lower face, suggesting that the change in morphology
resulted from decreased mechanical loading of the face and jaws brought about by consump-
tion of softer-textured foods in later prehistory (Boyd, 1988). Accompanying these changes is

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.2 Cranial form and functional adaptation 263

a marked reduction in size of the temporomandibular joint (Hinton, 1981a, 1983). Experi-
mental studies show that this joint is highly sensitive to alterations in mechanical loading
(Bouvier & Hylander, 1981, 1982, 1984; Tuominen et al., 1993), and the joint tends to be
largest in human populations with high masticatory stresses (Corruccini & Handler, 1980;
Hinton, 1981a, 1981b, 1981c, 1983; Hinton & Carlson, 1979; Wedel et al., 1978). Like the
assessments of robusticity and vault shape generally, these findings denote the primacy of
mechanical factors in interpreting craniofacial morphology, including when population
continuity can be established.
Temporal analysis of crania from a very similar setting from the Ohio River valley reveals
complementary evidence of masticatory function and craniofacial robusticity (Paschetta
et al., 2010). Comparison of foragers (Archaic-period Indian Knoll), incipient farmers
(Woodland-period sites, Kentucky), and intensive farmers (Mississippian-period sites) demon-
strates clear differences in craniofacial robusticity, with the temporal changes of reduction in
robusticity in the face (zygoma and alveolus reduction), and especially reduced projection of
the face, and reduced size of the temporomandibular joint. These changes link directly with
reduced masticatory loading of the temporalis muscles and increased consumption of softer
foods.
Similarly, comparisons of prehistoric Georgia coastal forager-fishers (pre-AD 1150) and
farmers (AD 1150–1550) reveal that, like the Tennessee and Ohio River valley populations,
there is a general decrease in craniofacial robusticity. Like these settings, the reductions
in facial and mandibular dimensions and attachment sites for masticatory muscles (tem-
poralis and masseter) were more pronounced than reduction in nonfacial dimensions
(Larsen, 1982, 1984). These changes in craniofacial size and robusticity in prehistoric
Georgia Indians appear to be due to increased consumption of soft, maize-based foods
in later prehistory (Larsen, 1982). These findings, therefore, are strongly suggestive of
responses of the craniofacial skeleton to change in subsistence and the manner in which
food is prepared.
Strikingly similar diachronic patterns of craniofacial robusticity are revealed in a suite of
analyses from South America in the comparison of earlier foragers and farmers/pastoralists
(González-José, Ramírez-Rozzi et al., 2005; Sardi & Béguelin, 2011; Sardi et al., 2006). In the
Argentine Center-West, the southernmost region of the Andes where the foraging-to-farming
transition took place, beginning c. 3000 BC and heavily emphasized by AD 600, the dietary
transition involved the adoption of beans, maize, potato, and manioc. These domesticated
plants were ground and boiled to a soft consistency. The differences in the comparison of the
earlier foragers and later farmers are profound, especially in cranial morphology and the
remarkable reduction in the masticatory apparatus in the farmers (Sardi & Béguelin, 2011;
Sardi et al., 2006). Like the transitions in North America, this regional change in South
America appears to represent a continent-wide trend in craniofacial modifications concomi-
tant with masticatory loading and differences associated with hunter-gatherers and
agriculturalists (González-José, Ramírez-Rozzi et al., 2005).
Thus, the record of evidence indicates that the worldwide trend of cranial shortening and
gracilization is best understood in relation to masticatory, dietary, and technological changes,
especially those associated with the shift from foraging to food production and consumption
of softer foods by later populations. These changes in subsistence practices and their influence
on craniofacial anatomy are global in scope (Larsen, 1995; von Cramon-Taubadel, 2011).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
264 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

While the general trends of reduction in craniofacial robusticity – especially in the mastica-
tory complex – are clear, it is important to emphasize that there is significant variation. For
example, in the southern Levant (Israel), the transition from foraging to farming had consider-
able effects on tooth size (reduction) but only modest effects on mandibular size (Pinhasi
et al., 2008). In this regard, analysis of a series of populations representing foragers and
farmers reveals a statistically significant reduction in ramus breadth and symphysis height
only. Although these changes likely represent responses to changing loading patterns, the
modest reductions reflect local circumstances in food preparation technology and the kinds of
foods consumed. Moreover, in South Africa, there is a reduction in craniofacial size and
robusticity (c. 2000–1000 BC) coincident with increased intensification and breadth of foods
consumed (Stynder et al., 2007). However, unlike other settings described earlier where this
kind of adaptive intensification led to domestication and farming (e.g., North America,
Europe, Levant), in this setting agriculture did not arise. Rather, the craniofacial changes in
South African foragers were due to some other adaptive shift, perhaps related to the economic
intensification unique to this particular setting. Finally, it is important to emphasize that
other, non-environmental factors play an important role in understanding craniofacial
changes. In the Great Plains region of North America, for example, morphometric analyses
implicate gene flow between neighboring populations rather than masticatory loading (Jantz,
1973; Key, 1983, 1994; Ubelaker & Jantz, 1986). This underscores the important point that
cranial morphology – as with all other phenotypic variation – reflects the interaction between
genes and environment.

7.2.3 The supraorbital torus: a beam?


The size and robusticity of the supraorbital torus has been investigated intensively by
biological anthropologists and anatomists for well over a century (see review by Russell,
1985). The development of the torus is highly variable in humans, ranging from a thick, bar-
like projection in early Homo (Weidenreich, 1943) to mild expression in most recent human
populations, with some notable exceptions (e.g., native populations in the American Great
Basin and Australia). Because only small muscles of facial expression are directly associated
with the supraorbital torus, the region is often interpreted as essentially nonfunctional and
nonadaptive (Owen, 1855; and later researchers). This is supported by feeding experiments on
laboratory monkeys whereby measurement of in vivo strains are quite low (Deagling, 2010;
Hylander et al., 1991), a finding consistent with finite element analysis of strain magnitude
(Kupczik et al., 2009). Simply, the evidence suggests that supraorbital tori size and morph-
ology are not related to mastication.
Although the size and morphology of the supraorbital torus are likely not related to
masticatory loading, the comparison of supraorbital development in temporal sequences of
recent humans from archaeological settings suggests that browridge size is best thought of as
a general indicator of overall craniofacial robusticity. That is, there is an allometric pattern
that involves the supraorbital torus: when other indicators of robusticity change, the supra-
orbital torus increases in its expression. Carlson (1976) found that the torus was more
developed in the preagricultural foragers than in farmers from Nubia (and see earlier). In
the American northern Great Plains (South Dakota and North Dakota), comparison of size and
morphology of the adult female and male supraorbital tori in Woodland tradition foragers

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.2 Cranial form and functional adaptation 265

(AD 610–1033) with mixed forager/agriculturalists from the Middle Missouri (AD 900–1675)
and Coalescent (AD 1600–1832) traditions reveals a general gracilization of the supraorbital
torus from the earlier to later periods (Cole & Cole, 1994).

7.2.4 Eskimo craniofacial morphology: masticatory loading or cold adaptation?


Craniofacial morphology of circumpolar groups – especially Eskimo populations – is charac-
terized by a pronounced degree of robusticity, including marked facial flatness, well-
developed and anteriorly placed zygomas, high and pronounced temporal lines, and extreme
overall robusticity of the face, jaws, and masticatory apparatus generally. Coon and cow-
orkers (1950) argued that Eskimo craniofacial morphology represents an adaptation to
extreme cold. For example, they interpreted the presence of enlarged and forwardly placed
zygomas as reflecting a retraction of the external nose, an area of the face that is especially
vulnerable to cold stress. Alternatively, others suggested that the Eskimo craniofacial morph-
ology represents adaptation to vigorous mastication (Furst & Hansen, 1915; Hrdlička, 1910b),
or what has been called the “hard-chewing” hypothesis (Collins, 1951).
In order to determine which of the two models best explains Eskimo craniofacial morph-
ology, Hylander (1977) undertook comprehensive biometrical and paleopathological analyses
of masticatory behavior in past and living Eskimos. His analysis reveals a link between
craniofacial morphology and loading of the jaws and teeth in these populations. For example,
many dentitions show root resorption and crown fractures and chipping due to excessive
mechanical demands. These populations also express high frequencies of mandibular, maxil-
lary, and palatine tori, skeletal features that have been linked with severe or elevated
masticatory stresses (see later). Bite force measured in living Alaskan Eskimos is remarkably
high compared to other, non-circumpolar, populations.
A rich body of ethnographic evidence indicates that very heavy mechanical demands are
placed on the craniofacial complex of Eskimos, specifically involving heavy use of jaws and
teeth in masticatory and extramasticatory functions. In his observations of Eskimos, De
Poncins (1941:71–72) noted, “They had long since stopped cutting the meat with their
circular knives; their teeth sufficed, and the bones of the seal cracked and splintered in
their faces. What those teeth could do, I already know. When the cover of a gasoline drum
could not be pried off with the fingers, an Eskimo would take it between his teeth and
it would come easily away. When a strap of sealskin freezes hard – and I know of nothing
tougher than sealskin – an Eskimo will put it in his mouth and chew it soft again” (quoted
in Hylander, 1977:142).
Drawing on these various lines of biological and behavioral evidence, Hylander (1977)
argued that Eskimo craniofacial morphology, long observed by biological anthropologists
(e.g., forwardly placed zygomas), is oriented toward maximizing the efficiency and power of
chewing, especially involving anterior tooth use: the craniofacial complex is suited to the
generation and dissipation of pronounced, vertically oriented masticatory forces in the front
of the mouth. This assessment is confirmed by the analysis of the position of attachment sites
for masseter and temporalis muscles and of the incisors in a sample of prehistoric Inuit crania
(Spencer & Demes, 1993). The anteriorly placed masticatory muscles and posteriorly placed
incisors “indicate an increased efficiency for the application of either high magnitude or
repeated bite forces on the anterior dentition” (Spencer & Demes, 1993:15).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
266 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

Other high latitude foragers display pronounced craniofacial robusticity. Like Eskimos,
crania from Tierra del Fuego and Patagonia, South America, bear robust supraorbital tori
and anteriorly placed zygomas, sagittal keeling, occipital tori, and pronounced attachment
sites for the temporalis muscle (Lahr, 1995). The functional-masticatory paradigm (Hylander,
1977; Spencer & Demes, 1993) best explains the similarities between Eskimos and Fueguian/
Patagonians, especially with regard to the common skeletal responses to highly demanding
masticatory regimes in these two different settings.
Some workers have argued that the round head of high latitude populations would be best
suited for cold adaptation, indicating that a sphere maximizes volume for heat retention and
minimizes surface area for heat loss prevention (Beals, 1972; Beals et al., 1984; Crognier,
1981). However, the aforementioned studies linking craniofacial morphology with trends in
masticatory loading and dietary change in the Holocene and changes in head shape in the
absence of climatic change makes thermoregulatory models or similar types of ecological
arguments less than compelling (Armstrong, 1984; Gibson, 1984; Henke, 1984; Henneberg,
1984; Lahr, 1995; Morimoto, 1984). In North America, analysis of cranial dimensions based
on thousands of native individuals measured by Boas in the late nineteenth century shows no
relationship between climate and head shape (Jantz et al., 1992; and see De Azevedo et al.,
2011). Finally, the cold stress model has limited explanatory power because a wide range of
other human populations with round heads live in warm climates, and various populations,
living and extinct, having robust, forwardly placed molars are associated with hot, dry
climates (Hylander, 1977). Moreover, recent global analyses show that climatic adaptation
does not explain cranial shape differences at a global level but that it might explain thermo-
regulatory adaptation in high-latitude populations (Harvati & Weaver, 2006; Hubbe et al.,
2009; Noback et al., 2011; von Cramon-Taubadel, 2009a, 2011).

7.2.5 Incisor shoveling and masticatory loading


Incisor shoveling has been observed in a wide range of human populations worldwide, but the
highest frequencies of the well-developed form appear to be found in high-latitude and cold-
adapted foragers (Mizoguchi, 1985). Mizoguchi (1985) argues that meat eating in these
hunter-gatherers requires heavier masticatory loading of the anterior dentition than do other
foods in most other regions of the world (e.g., pastoralists in Africa). His assessment suggests a
possible link between incisor morphology and demands of pronounced incisor loading.

7.2.6 Palatine, maxillary, and mandibular tori and masticatory stress


Oral hard tissue tori, mostly located on the hard palate and lingual corpus and alveoli of the
mandible, are the focus of attention regarding genetic versus nongenetic environmental
influences on skeletal variation (Halffman et al., 1992; Hauser & De Stefano, 1989; Morris,
1981; Sirirungrojying & Kerdpon, 1999) (Figure 7.4). Some workers conclude that tori found
in archaeological remains are indicative of high mechanical demands placed on the mastica-
tory apparatus (Halffman et al., 1992; Hooton, 1918; Hrdlička, 1940b; Pechenkina et al., 2002;
Pedersen, 1944; Scott et al., 1992), but others argue that tori are largely genetically controlled
(Gorsky et al., 1998; Hauser & De Stefano, 1989; Morris, 1981).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.2 Cranial form and functional adaptation 267

Figure 7.4 Superior (top) view of large mandibular tori: Kodiak Island, Larsen Bay, Alaska,
pre-AD 1000. (Photograph by G. Richard Scott.) (A black and white version of this figure
will appear in some formats. For the color version, please refer to the plate section.)

There are relatively high frequencies of palatine tori in far northern and circumpolar
populations, including Icelanders (Hooton, 1918), Lapplanders (Schreinder, 1935), and Eskimos
(Hylander, 1977; and prior discussion). Torus prevalence is also high in Medieval Norse living in
Iceland and Greenland in comparison with Europeans generally (Mellquist & Sandberg, 1939;
Pedersen, 1944). In order to address the question of why Medieval Norse and indigenous Arctic
groups have developed a convergence in torus expression with Eskimos and other circumpolar
groups, Halffman and coworkers (1992; Scott et al., 1992) studied torus frequency and size in a
series of Medieval Norse skeletal remains from Norway, Greenland, and Iceland dating to the
eleventh through fourteenth centuries. Temporal comparisons reveal that later Norse from
Norway, Greenland, and Iceland have a significantly higher prevalence of tori than do early
Norse from Greenland. Tori prevalence approaches 100% in later populations, which is among
the highest in the world (compare with Halffman et al., 1992:156; Hauser & DeStefano, 1989).
Tori increase in size in older adults, suggesting that the trait is strongly influenced by age. The
pattern of increase in tori frequency and size within a relatively short temporal span (several
hundred years) suggests that environment in general and masticatory loading in particular
exert more influence on torus expression than does inheritance.
Changes in subsistence and food preparation techniques provide some insight into possible
reasons for the secular increase in tori in Greenland. Archaeological evidence indicates that
the subsistence economy of later Greenlandic Norse became increasingly focused on wild
game (e.g., seals and caribou), rather than domestic animals (cattle, sheep, and goats). Easily
chewed foods such as grains and breads became increasingly scarce in the later Norse in
Greenland. Archaeological evidence indicates a decrease in cooking of food due to declines in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
268 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

availability of firewood. As such, this would have involved an increase in consumption of


tougher foods, namely raw or partially cooked foods, including meat. Scott and coworkers
(1992) document a general increase in tooth wear and anterior tooth crown chipping in the
later period. These findings are consistent with the hypothesis that masticatory function is the
principal influence on tori expression.
The clear links between high levels masticatory stress, as it is represented in advanced
stages of occlusal surface wear in a number of settings, speaks to the important role of
subsistence practices in tori expression. For example, tori prevalence in various archaeological
contexts in northern and central China shows a strong association with advanced tooth wear
(Liu et al., 2010; Pechenkina, Benfer et al., 2002, 2007). Temporal comparisons of early and
late Neolithic populations in Shaanxi Province show a marked decline in tori prevalence and
tooth wear as farming intensified and foods consumed became softer. In these contexts, tori
prevalence increased with age, reflecting the cumulative nature of masticatory loading over
the life course (Pechenkina et al., 2002). Similarly, comparisons of Bronze and Iron Age
hunter-gatherers in northwest China (Xinjiang Province) with Neolithic farmers in central
China reveal much higher prevalence and severity of tori in the former than in the latter (Liu
et al., 2010).

7.2.7 Pathological modifications of the temporomandibular joint


Like the other joints of the skeleton, the temporomandibular joint is subject to mechanical
demands leading to osteoarthritis. Populations experiencing high masticatory loading show a
tendency for elevated prevalences of temporomandibular joint degenerative pathology
(Brown, 1992; Richards & Brown, 1981; Webb, 1995; Wells, 1975). Inuit have an unusually
high prevalence of temporomandibular joint osteoarthritis, with women expressing higher
frequencies than men (Merbs, 1983). These differences may be related to the preparation of
hides with the teeth, a task performed primarily by women in Eskimo societies (Merbs, 1983).
Sedentary agriculturalists from the Medieval period Kulubnarti site in northern Sudan have
similarly high levels of articular pathology, with females having three times that of males
(Sheridan et al., 1991). In general, human populations with elevated prevalences of temporo-
mandibular joint osteoarthritis also possess heavy occlusal wear, thus indicating the strong
influence of mechanical stress in temporomandibular joint degeneration (Brown, 1992; Liu
et al., 2010; Pechenkina et al., 2002; Webb, 1995; Wells, 1975; Whittaker, 1993). Moreover, in
temporal comparisons of high levels of masticatory loading followed by decline in mastica-
tory loading in the context of increased intensification of farming, there is a clear decline in
osteoarthritis of this joint, from 52% to 25% (Pechenkina, Benfer et al., 2002, 2007).

7.2.8 Age changes in craniofacial size and morphology in adulthood


As discussed earlier, the development of skeletal form is a dynamic process that begins well
before birth and continues in the months and years of growth and development (Albert et al.,
2007; Behrents, 1985; Holmes & Ruff, 2011). An important implication is that growth in the
skull is not quiescent once adulthood is reached. Like the postcranial skeleton, appositional
bone growth in the skull continues during adulthood through remodeling as part of the
normal aging process. Throughout the lifespan, the craniofacial skeleton increases cranial

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.2 Cranial form and functional adaptation 269

vault circumference, length, bizygomatic breadth, and head breadth. One of the most dynamic
areas of size increase is height of the lower face, including the mandible, with relatively little
change in the upper face (Forsberg et al., 1991). Interestingly, this is also the most dynamic
area of adaptation in masticatory stress as demonstrated by experimental and population
studies (see prior discussion). The results of this process are documented in cross-sectional
studies involving comparisons of individuals of different ages within the same population
(Baer, 1956; Goldstein, 1936; Hooton & Dupertuis, 1951; Howells & Bleibtreu, 1970; Hrdlička,
1936; Lasker, 1953; Todd, 1924; Walker, 1995; and see Albert et al., 2007), and in longitudinal
studies comparing sequential observations of the same individual at different ages (Behrents,
1985; Israel, 1968, 1973, 1977; but see Tallgren, 1974). Overall, these investigations show the
dynamic nature of craniofacial architecture during adulthood, especially with regard to
increases in craniofacial dimensions, even in later years when muscle mass reduces (Albert
et al., 2007; Garn, 1985).
Most investigations of bone apposition are based on industrialized urban populations, and
thus represent a narrow perspective on human variation. Studies of archaeological samples
are important in that they provide a wider context for assessing the roles of genes and
environment. In this regard, Ruff (1980) compared a large sample (n¼136) of adult males
from the Archaic-period Indian Knoll, Kentucky series. Statistical comparisons of younger
(20–34 years of age) with older (35–50 years of age) males revealed that virtually all
dimensions (e.g., face width, face height) were larger in the older adults, six of which reached
significance (P<0.05). Comparisons of craniofacial dimensions in a series of adult females and
males from Late Woodland and Mississippian-period sites in west-central Illinois (Droessler,
1981) revealed similar trends to the Indian Knoll series. Comparisons of young adults (20–34
years old), middle-aged adults (aged 35–49 years), and old adults (50þ years of age) indicate
that cranial size tends to increase with age, but to a relatively greater extent in the face, and
more so in females than in males (Table 7.1). Age comparisons of crania from Europe,
Melanesia, and the American Great Plains also revealed significant increases in adult cranio-
facial size (Guagliardo, 1982b; Sejrsen et al., 1996). Increases in facial projection, interorbital
width, orbit and mastoid size (Guagliardo, 1982b), and palate width (Sejrsen et al., 1996) are
especially prominent.
Although these studies show generally increasing size with advancing age, there is some
intra-population variability, such as changes in females but not in males (Albert et al., 2007;
Droessler, 1981), earlier changes in adult males than in adult females (Walker, 1995), or lack
of age changes entirely (Lynnerup, 2001; Sjøvold, 1995). This variability, especially between
males and females, may reflect differences in tooth loss or some unknown factors (Droessler,
1981). The congruence of results from studies of populations representing very diverse
lifestyles and ecological settings suggests some common factor or factors – genetic and/or
environmental – that influence bone apposition during adulthood.
Baer (1956) suggested three possible explanations: secular change, selective survival, and
true ontogenetic development. In working with archaeological samples, special problems
unique to these kinds of data sets emerge. For example, archaeological series usually include
multiple generations spanning many years and differing adaptive regimes. The west-central
Illinois series, for example, encompasses about 600 years of occupation with a major dietary
shift from foraging to farming (Droessler, 1981). The consistency of craniofacial changes with
advancing age in investigations of modern (clinical) and archaeological populations is

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
270 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

Table 7.1 Age comparisons of west-central Illinois adult cranial measurements: young (20–34
years), middle (35–49 years), old (50þ years) individuals (Adapted from Droessler, 1981. All
measurements in millimeters)
Young (n¼54)a Middle (n¼58)a Old (n¼39)a

Mean SD Mean SD Mean SD


Measurement
Males
Vault length 180.9 6.3 182.3 7.1 180.4 6.0
Vault breadth 138.5 5.6 137.1 5.8 138.5 4.5
Vault height 142.3 5.4 140.3 5.3 140.4 4.3
Face height 122.9 5.8 122.1 6.8 122.1 4.2
Mid-face breadth 99.8 5.2 99.9 4.8 100.3 4.8
Females
Vault length 173.0 5.9 173.6 5.1 174.2 6.7
Vault breadth 134.0 4.3 134.7 4.5 135.1 6.6
Vault height 136.5 4.2 137.5 5.2 137.3 4.8
Face height 114.5 6.0 112.6 5.8 114.4 7.2
Mid-face breadth 96.4 4.2 95.5 5.2 97.2 4.4
a
Average sample size; sample size varies by measurement.

striking, suggesting that the trends are more real than apparent. As discussed previously,
biomechanical factors likely influence craniofacial expansion, including changes with age
due to the cumulative forces associated with mastication.
Regardless of cause, the increasing expansion of the craniofacial complex with advancing
age underscores the importance of considering the demographic composition of skeletal and
clinical series when making comparisons and drawing inferences about craniofacial morph-
ology. Comparison of a predominantly young adult series from one population with an old
adult series to another might give the impression of differences in robusticity. Consideration
of age structure, therefore, represents a vital part of skeletal analysis. These craniofacial size
differences are most apparent in the comparison between very young and very old adults
(Behrents, 1985). Most archaeological samples contain relatively few older adults. Thus,
differences in craniofacial robusticity between different archaeological series are likely due
to factors other than age.

7.3 Dental and alveolar changes


7.3.1 Occlusal abnormalities and dental crowding
Occlusal abnormalities and dental crowding are generally lumped under the term “malocclu-
sion,” which includes all manner of conditions deviating from the “ideal” or normal occlusion.
Angle’s (1898) highly influential treatise on malocclusion defined ideal occlusion as “edge-to-
edge” whereby all antimeric teeth in the upper and lower dentitions are in perfect occlusion.
This pattern is exceedingly rare, especially in Western industrialized populations or
populations adopting Western diets and subsistence technology. Abnormal occlusal
patterns in humans fall into two broad categories: (1) dental crowding involving insufficient

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.3 Dental and alveolar changes 271

space for all teeth (e.g., tooth impaction), and (2) underdevelopment or overdevelopment of
the maxillary dentition relative to the mandibular dentition or vice versa (e.g., overbite and
overjet) (Corruccini, 1991; Proffit et al., 2013; Scott & Turner, 1988).
The causes of malocclusion have been debated for years, some workers invoking genetic
explanations (Hrdlička, 1922b; Lundström, 1948; Smith & Bailit, 1977) and others offering
environmental explanations (Begg, 1954; Hunt, 1960, 1961; Sagne, 1976). Comparisons of
monkeys (squirrel monkeys, macaques, and baboons) fed hard and soft diets reveal significant
differences in occlusion; animals fed soft diets tend to have narrower maxillary arches, higher
frequencies of tooth impactions, rotations, and crowding than do animals fed hard diets
(Beecher & Corruccini, 1981; Corruccini and Beecher, 1982, 1984). Human studies are consist-
ent with these findings. In Western societies where highly processed foods make up a consider-
able part of diet, the prevalence of malocclusions exceeds half of the population (Evensen &
Øgaard, 2007; and see Proffit et al., 2013). In contrast, non-Western societies consuming
traditional diets composed of hard-textured foods have a very low prevalence of occlusal
abnormalities (Corruccini et al., 1981, 1983; Hunt, 1961; Lombardi & Bailit, 1972; Lu, 1977;
Moorrees, 1957; Price, 1936). In addition, Corruccini and Whitley (1981) showed that older
Euroamericans in rural Kentucky who had been raised on traditional diets (e.g., dried pork,
heavy corn-bread, wild garden foods) have a low prevalence of malocclusion. In contrast, high-
status South Asians who have greater access to soft, refined foods than low-status South Asians
have an elevated prevalence of occlusal abnormalities (Corruccini et al., 1983). Similarly, urban
Greek populations consuming nontraditional diets dominated by soft, processed foods have
more dental crowding and non-edge-to-edge bite than rural Greek populations consuming
traditional diets (Angel, 1944). These changes reflect the secular trend of increased malocclu-
sions (Corruccini, 1984). Similarly, comparisons of Medieval Norwegians with modern
Norwegians reveal a relatively low prevalence of malocclusions (35%) in the former compared
to the latter (65%), with females showing greater prevalence than males in the Medieval series
but no difference in modern Norwegians (Eversen & Øgaard, 2007).
Populations shifting to highly processed, nontraditional diets show an increase in overbite,
overjet, impactions, crowding (Brown, 1992; Corruccini, 1984; Corruccini & Choudhury,
1985; Corruccini et al., 1981, 1990; Moorrees, 1957; Price, 1936; Waugh, 1937; Wood,
1971), and a general narrowing of the arches (Corruccini & Whitley, 1981; Goose, 1962,
1972; Lundström & Lysell, 1953; Lysell, 1958). The latter findings suggest that the
narrower faces of American-born children in comparison with their European-born parents
(Boas, 1912, 1916; Hrdlička, cited in Boas, 1916) resulted from the shift to softer, more
processed foods.
The association between malocclusion and consumption of soft-textured foods in animal
and human studies supports the “disuse” hypothesis; namely, reduction in masticatory
mechanical loading has led to a reduction in growth of bone tissue supporting the teeth.
Certainly, some populations appear to have a genetic predisposition for occlusal abnormalities
(Corruccini, 1991), and there is definitely a genetic component to dentofacial abnormalities in
general (Carlson, 2005). However, overwhelming evidence indicates that reduction in masti-
catory stress has engendered an elevation in occlusal abnormalities in recent human history.
From this, it can be implied that tooth size has reduced over the course of human evolution
(and see later) at a relatively slower rate than that of the supporting bony structures of the
maxilla and mandible. Differential reduction in tooth and jaw size has been documented in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
272 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

recent human populations. Comparisons of British orofacial and dental dimensions spanning
the Neolithic to the nineteenth century show a distinctive reduction in bone dimensions (e.g.,
mandible, palate width) but no or little reduction in tooth size (Goose, 1963; Keith, 1924;
Moore et al., 1968). Similarly, comparisons of Medieval and recent Swedish skulls show
reduction in mandible size but not in tooth size (Sagne, 1976). These findings indicate that
temporal increases in crowding are essentially reductions in bone size without corresponding
decreases in tooth size. This differential reduction is due to the fact that, unlike bone, teeth
cannot respond to changes in use via differential growth. Thus, while jaw size reduces in
response to dietary changes, tooth size is not able to do so.
Findings based on temporal comparisons from several archaeological contexts are consist-
ent with the disuse hypothesis concerning the replacement of hard-textured by soft-textured
foods in the foraging-to-farming transition or within increasing urbanization. Comparison of
Archaic hunter-gatherers with Mississippian agriculturalists from Koger’s Island in the Pick-
wick Basin, Tennessee, reveals an increase in occlusal abnormalities in the agriculturalists
relative to the hunter-gatherers (Newman & Snow, 1942). Because there was also a decline in
tooth wear in this setting, Newman and Snow (1942) concluded that dietary consistency
declined due to the use of wooden mortars and pestles in food preparation in later prehistory,
thus indirectly implicating decreased masticatory stress as a factor contributing to increased
prevalence of malocclusion.
In the eastern Mediterranean, decrease in edge-to-edge bite from the Neolithic to the mid-
twentieth century also coincides with a reduction in tooth wear, which led Angel (1944) to
infer a decrease in mechanical demand on the faces and jaws. The pattern was also linked to a
decline in craniofacial robusticity, accelerating dramatically in recent Greek populations.
The trend of increasing prevalence of occlusal abnormalities has been documented in a
variety of settings and interpreted in various ways (compare with Davies, 1972; Dickson,
1970). This trend is well documented in archaeological remains from Japan, especially in view
of changes in diet, food preparation technology, and orofacial adaptation (Hanihara et al.,
1981; Inoue et al., 1986; Ito et al., 1983; Kaifu, 1999, 2000; Kamegai et al., 1982). Compari-
sons of adult skulls from the Jomon (1000–500 BC), Yayoi, Kofun (AD 50–350), Medieval
(AD 1300–1600), Yedo (AD 1600–1900), and modern (AD 1964–1966) periods reveal clear
trends in orofacial adaptation (Hanihara et al., 1981; Kaifu, 1999, 2000). Frequency of
malocclusion increased from 20.0% in Jomon hunter-gatherers to 45.5% in Kofun incipient
agriculturalists, to 76.2% in modern Japanese. Building on these findings, Inoue and
coworkers examined the relationship between food consistency and occlusal abnormalities
in Japan, succinctly arguing that “soft and highly nutritious food reduces the functional
activity of the human masticating system, because it requires less chewing force and less
chewing time. Thus reduction of the jaw bone has progressed through the course of human
micro-evolution, resulting in disharmony between the sizes of teeth and jaws” (Inoue et al.,
1986:164; and see Suzuki, 1969). To test this hypothesis further, malocclusion prevalences of
Jomon-period foragers (divided into an early and late sample), Yayoi-period early farmers,
and Kofun-period protohistoric farmers mostly from southern Japan were compared. Factor,
principal components, and cluster statistical analyses indicated a grouping of occlusal pat-
terns based on subsistence categories. This analysis revealed a low prevalence of malocclu-
sions in the foragers followed by increases in the later, agricultural populations with increased
use of soft foods. Among the most striking temporal changes in this complex of craniofacial

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.3 Dental and alveolar changes 273

changes is protrusion of the anterior dentition, overbite, and increasing alveolar prognathism
in later Japanese relative to earlier populations (Kaifu, 1999, 2000). Almost certainly, highly
distinctive changes are related to decreasing masticatory loading, perhaps relating to the
greater reduction of supporting bony architecture relative to large anterior teeth. However, the
role of anterior tooth wear – especially its considerable reduction and the associated conse-
quences for increasing malocclusion – is an essential component for understanding these very
significant craniofacial changes in East Asia and elsewhere.
Collectively, the record of skeletal, tooth size, and tooth wear research reveals a complex
interaction of bone, tooth size, and rate of dental attrition. It may be the case that a number of
epigenetic mechanisms are responsible for the integration of craniofacial growth, tooth
growth, and tooth wear (Lieberman, 2011). While it is the case that tooth size has a higher
genetic component than the masticatory skeletal apparatus, permanent tooth size nonetheless
has a significant environmental component (Harris & Rathbun, 1991; and see discussion in
Lieberman, 2011). Tooth size has decreased over the last 1000 years (Brace et al., 1991).
Moreover, in China, posterior tooth size has increased in just the last several generations
relating to improved nutrition and increased body size (Harris et al., 2001). Lieberman (2011)
speculates that just as chewing of hard food stimulates jaws and tooth roots to become larger,
such a stimulus may also act on tooth crowns during their development to become larger.

7.3.2 Tooth size changes


The adaptational and evolutionary significance of past tooth size variation in human popula-
tions is a major point of discussion in bioarchaeology, as is indicated by a rich body of
research on the topic from archaeological settings in the Old World (Brace & Hinton, 1981;
Brace et al., 1991; Calcagno, 1989; Calcagno & Gibson, 1991; Jacobs, 1994; Lukacs, 1985;
Lukacs & Hemphill, 1991; Pinhasi et al., 2008; Smith, Bar-Yosef et al., 1984; Walimbe &
Kulkarni, 1993; y’Edynak, 1989) and in the New World (Brace & Mahler, 1971; Dahlberg,
1963; Hinton et al., 1980; Larsen, 1982; Sciulli, 1979; Simpson et al., 1990; Walker, 1978).
These studies emphasize the central importance of teeth in the understanding of craniofacial
adaptation and dietary change throughout the hominin evolution, including the appearance
and evolution of recent Homo sapiens. This is especially so because these changes occurred in
a short amount of time, at least from an evolutionary perspective (Hillson, 2005).
The trend for a reduction in tooth size – locally, regionally, and globally – over the course of
hominin evolution has been well documented (Bermúdez de Castro & Nicolas, 1995; Brace,
1995; Brace & Mahler, 1971; Brace et al., 1987, 1991; Calcagno, 1986a, 1986b, 1989;
Calcagno & Gibson, 1991; Eshed et al., 2006; Frayer, 1978; Harris & Lease, 2005; Kieser,
1990; Larsen, 1982, 1983; Lukacs, 1985; Pinhasi et al., 2008; Scott & Turner, 1988; y’Edynak,
1989; but see Jacobs, 1994; Scott, 1979) (Figure 7.5). Generally, foragers or groups having
recently made the shift to agriculture have larger teeth than populations with a longer history
of agricultural use and associated food preparation techniques, such as boiling in ceramic
vessels and other forms of extended cooking (Brace & Hinton, 1981; Brace & Mahler, 1971;
Brace et al., 1987, 1991; Lukacs, 1985). In the Levant (Israel), detailed analysis of trends with
regard to the incremental transition from the beginning of a sedentary lifestyle (Natufian)
through the adoption and intensification of farming (Neolithic) reveals significant reductions
in tooth size, especially in buccolingual diameters (Pinhasi et al., 2008). This trend in tooth

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
274 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

(a)
1600

Summed Posterior Areas (mm2) 1500

1400

1300

1200

1100

1000

900

800
al . .
lis us od od
bi ec
t r th m m
. ha er de rly te
. n la
H H ea ea
N
Hominid
Figure 7.5a Reduction in summed posterior tooth size in Homo from early species to
anatomically modern humans. (Data from Calcagno & Gibson, 1991.)

breadth is consistent with other analyses of Holocene dental size reduction in populations
undergoing the shift from foraging to farming in Europe (Brace et al., 1987; Pinhasi &
Mieklejohn, 2011) and northeast Africa (Calcagno, 1989).
Yet, mechanisms for tooth size reduction remain elusive (see Kieser, 1990, and Pinhasi
et al., 2008 for review of alternative models explaining tooth size reduction). Tooth size
appears to be highly heritable, suggesting that size reduction may be largely an evolutionary
(genetic) change. The rapid reduction in tooth size accompanying shifts in dietary focus and
other circumstances (Harris et al., 2001; Hinton et al., 1980; Larsen, 1982, 1983; Sciulli, 1979)
or the presence of small teeth in physiologically stressed individuals (Garn & Burdi, 1971;
Garn, Osborne et al., 1979; Kieser, 1990; Townsend & Brown, 1978) underscores the influence
of environment (e.g., nutrition, maternal health). The greater heritability of tooth size than
bone size and the fact that teeth do not remodel once formed underlies the potential imbal-
ance between teeth and supporting bone, as is so well illustrated in the aforementioned studies
of occlusal abnormalities.
Large teeth located in proportionately small alveolar tissue also represent a potential health
risk, predisposing an individual to impactions, dental caries, periodontal disease, and tooth
loss (Calcagno & Gibson, 1991; Inoue et al., 1986). Caries and dental impactions can also lead
to abscesses and localized infections, which can progress to more severe types of systemic
infections (e.g., gangrene, septicemia, osteomyelitis; Calcagno & Gibson, 1991). Because

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.3 Dental and alveolar changes 275

(b)
UP3BL
12.5
UM2BL
12 LP4BL
11.5

11
Tooth length (mm)

10.5

10

9.5

8.5

7.5
Natufian

Natufian

Natufian

PPNA

PPNB

PPNC
Middle
Early

Final

Period
Figure 7.5b Reduction in upper premolar (UP3BL), upper second molar (UM2BL), and lower
premolar (LP4BL) lengths from Early Natufian (preagricultural) to Pre-pottery Neolithic C
(PPNC) periods (agricultural). (Data from Pinhasi et al., 2008.)

(c)
14
14
12 Male mandibular teeth Male maxillary teeth
12
10
Breadth (mm)

Breadth (mm)

10
8
8
6 6
Final Paleolithic Final Paleolithic

4 Agriculturalist Agriculturalist
4
Intensive Agriculturalist Intensive Agriculturalist
2 2
0 0
I1 I2 C P3 P4 M1 M2 M3 I1 I2 C P3 P4 M1 M2 M3

12 14
Female mandibular teeth 12 Female maxillary teeth
10
Breadth (mm)

10
Breadth (mm)

8
8
6
6
4 Final Paleolithic Final Paleolithic
Agriculturalist 4 Agriculturalist
2 Intensive Agriculturalist
2 Intensive Agriculturalist

0 0
I1 I2 C P3 P4 M1 M2 M3 I1 I2 C P3 P4 M1 M2 M3

Figure 7.5c Significant reduction in breadth of all teeth between Nubian Final Paleolithic
(preagriculturist) and agriculturalist populations. (Data from Calcagno, 1986a.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
276 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

(d) 140
EUP

120 LUP

MESO
100
Vexin Medieval
Tooth area (mm2)

80

60

40

20

0
C P3 P4 M1 M2 M3 C P3 P4 M1 M2 M3

Mandible Maxilla

Teeth

Figure 7.5d Overall decreases in mandibular and maxillary tooth areas of populations from
Early Upper (EUP) and Late Upper (LUP) Paleolithic periods to Mesolithic (MESO) and
Merovigian (Vexin Medieval) periods. (Data from Frayer, 1977.)

systemic infections are potentially life-threatening, teeth in populations consuming soft,


cariogenic foods may be under selection for reduction in these circumstances (Calcagno &
Gibson, 1991). However, small teeth in highly abrasive masticatory environments may wear
too rapidly, thus resulting in premature loss of crown height and pulp exposure. Exposed pulp
is highly susceptible to bacterial infection (pulpitis). This suggests, then, that teeth may also be
under selection for large occlusal size in abrasive environments.
The previous discussion highlights the extreme ends of variation involving the costs
and consequences of soft and hard diets in relation to mastication, tooth size, and potential
selective conditions. Most human populations lie somewhere between these ends of the spec-
trum of masticatory adaptation. Tooth size, therefore, is most likely a product of an ongoing
“selective compromise” involving the promotion of optimum dental health, among other things
(Calcagno & Gibson, 1991; Pinhasi et al., 2008). Several other models of tooth size reduction are
based on the premise that reduction occurs in the absence of selection (compare with Brace et al.,
1991; Macchiarelli & Bondioli, 1986), but these models exclude a broader consideration of
the potential health consequences of occlusal abnormalities. Thus, it appears that there is an
adaptive advantage to maintaining a harmonious relationship between teeth and the skeletal
structures supporting them (for an alternative perspective, see Kieser et al., 1985).

7.4 Dental wear and function


Use of the teeth in eating involves a two-stage process: first, the initial preparation of food with
the anterior teeth (incisors and canines), and second, the reduction of food with the posterior
teeth (premolars and molars). These activities result in wearing of surfaces as the upper and
lower teeth come into contact with each other, with the food being prepared or reduced, and

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 277

with grit inclusions within the food. Wear can also occur on adjacent surfaces between teeth as
well as food movement against the cheeks and tongue, which causes a relatively small amount
of wear, enough to obliterate perikymata, for example (Lucas, 2004). Over the course of human
evolution, there has been a shift away from the use of the teeth toward the use of the hands in
the manipulation of the environment and a reduction in the importance of mastication in the
consumption of processed foods. It is also true, however, that the presence of significant
amounts and variable patterns of occlusal surface wear, especially, in past and living human
populations, indicates the continued importance of teeth in survival and adaptation.
Two forms of wear, abrasion and attrition, are most commonly discussed (Kaidonis et al.,
1993; Lucas, 2004; Townsend et al., 1994). Abrasion is caused by contact between the tooth
and food or other solid exogenous materials, especially as food is forced over occlusal
surfaces. Attrition is caused by tooth-on-tooth contact in the absence of food or various
other abrasives. Additionally, erosion – the loss of tooth surfaces due to chemical dissolution –
is sometimes considered as a form of wear (Davis & Winter, 1980; Linkosalo & Markkanen,
1985). Because of the difficulty of distinguishing between abrasion, attrition, and erosion, this
discussion regards “wear” as any combination of abrasion, attrition, and erosion.
Tooth wear shows considerable variation between human populations. Owing to localized
behavioral characteristics, and differences in cultural practices, age, sex, diet, and orofacial
morphology, it provides enormously important information on earlier foodways and masti-
catory behavior (Benfer & Edwards, 1991; Clement & Hillson, 2012; Hillson, 2005; Molnar,
1971, 1972; Molnar & Molnar, 1990; Molnar et al., 1983; Powell, 1985; Richards, 1990;
Richards & Miller, 1991; Rose & Ungar, 1998; Walker et al., 1991). The significance of dental
wear in relation to diet was succinctly summarized by Walker, who noted that “[f]rom an
archaeological standpoint, dietary information based on the analysis of [wear] is of consider-
able value since it offers an independent check against reconstruction of prehistoric subsist-
ence based on the analysis of floral, faunal and artifactual evidence” 61978:101).
Dental wear variation is largely described and analyzed according to its severity or form. In
addition to occlusal surface wear, wear can involve substantial losses in the regions of contact
between teeth (Begg, 1954; van Reenen & Reinach, 1988; Wolpoff, 1971), resulting in
reduction in length (mesiodistal). Although severe wear can predispose a tooth to disease
(e.g., pulpitis, caries) and loss, tooth wear is a normal physiological process showing predict-
able modifications rather than a disease or pathological state. Viewed from this perspective,
tooth wear is a functional adaptation to stresses placed on teeth during mastication.
The physiological process of tooth wear is well understood. With regard to the most
common type of wear, occlusal surface wear, it commences with loss of occlusal enamel
and often the loss of cusps and crown height. If dentin is exposed, there is deposition of
secondary dentin serving as a protective zone overlying the pulp chamber. In extreme forms
of wear, all enamel is removed from the occlusal surface, leaving a nearly continuous surface
of dentin surrounded by a partial or complete rim of enamel. As the process continues in older
adults, the nerve supply in the pulp chamber withdraws toward the root tip and is replaced in
the pulp chamber by dentin. This replacement process is essential because it accommodates
extreme wear – in some populations, involving complete loss of crown height (Hartnady &
Rose, 1991; Larsen & Kelly, 1995). Moderate to severe wear results in a reduction in tooth size
due to the combined effects of occlusal and interproximal wear (van Reenen, 1982). In
especially severe wear environments, loss of crown size in molars appears to be compensated

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
278 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

Figure 7.6 Lingual tilting of mandibular first molar; Golovin Bay, Alaska, protohistoric.
(Photograph by G. Richard Scott.) (A black and white version of this figure will appear in
some formats. For the color version, please refer to the plate section.)

for by a combination of extra-eruption and lingual tilting (Clarke & Hirsch, 1991; Comuzzie &
Steele, 1989; Reinhardt, 1983; Taylor, 1963, 1986a, 1986b) (Figure 7.6). The added chewing
surface on the buccal aspects of the molar roots in these teeth likely represents an adaptive
response to severe wear by maintaining occlusal surface area commensurate with the dietary
or masticatory needs of the individual.
Severity of wear is highly influenced by the consistency and texture of food, which is
determined by either the characteristics of the food (e.g., presence of phytoliths or cellulose in
plants), the manner of its preparation, or some combination of these. In order to facilitate the
ready availability of nutrients to the digestive enzymes or to remove undesirable constituents
(e.g., fiber, toxins), humans process plants in a variety of ways, including grinding, pounding,
grating, soaking, leaching, drying, heating, and fermenting (Stahl, 1989). Some of these
processing techniques may involve the introduction of abrasive elements that promote tooth
wear (e.g., use of grinding stones for making flour from cereal grains), or they may involve the
removal of abrasives as in highly processed foods consumed by Western industrialized
populations. The following sections discuss wear as documented by two primary means,
namely through visual inspection of gross or macrowear and observation of microwear.

7.4.1 Macrowear
Some of the most comprehensive information on the behavioral significance of tooth wear is
from temporally successive series of human remains representing populations that had
undergone shifts in diet. Investigation of tooth wear and the mechanical environment reveals

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 279

a high degree of consistency with skeletal indicators of masticatory stress. Populations with
high levels of mechanical demand and/or reliance on abrasive diets have relatively advanced
wear (Barondess & Sauer, 1985; Clement & Hillson, 2012; Hansen et al., 1991; Hartnady &
Rose, 1991; Hemphill, 1992; Marks et al., 1988; Molnar et al., 1983; Powell & Steele, 1994;
Scott et al., 1992; Walker, 1978). Traditional populations undergoing shifts from diets con-
taining tough foods to diets dominated by processed foods during the twentieth century show
reduction in wear (Davies & Pedersen, 1955; Staz, 1938). Important differences in interprox-
imal wear and specific patterns of wear also help inform our understanding of masticatory
behavior in human populations.

Occlusal surface wear severity


Temporal comparisons of populations undergoing subsistence change within regions, especially
with regard to comparison of dentitions of foragers and later farmers, demonstrate the important
links between diet, masticatory behavior, and wear. A wide variety of human populations show
appreciable declines in severity of occlusal surface wear that accompany this shift (Larsen, 1995)
or in comparison of foragers and farmers from different settings in general (Deter, 2009).
Although various factors are no doubt involved in explaining this highly consistent pattern,
the common theme in explaining this dichotomy is the fact that the shift from foraging to
farming involved a change from eating generally more hard-textured (and sometimes highly
abrasive) foods to soft-textured foods. In the prehistoric Eastern Woodlands and other areas of
North America, comparisons of hunter-gatherers with later agriculturalists show a consistent
pattern of decline in severity of wear, which has been documented in the Tennessee River valley
(Hinton et al., 1980; Newman & Snow, 1942; Smith, 1982), the Roanoke River Valley in Virginia
(Hoyme & Bass, 1962), the Lower Mississippi River valley (Rose, Marks et al., 1991), the southern
border region of Oklahoma and Arkansas (Powell, 1985), western Pennsylvania and Ohio River
valley (Sciulli, 1997; Sciulli & Carlisle, 1977), Ontario (Patterson, 1984), the Tehuacan Valley of
Mexico (Anderson, 1965, 1967), and northern Chile (Alfonso et al., 2007). In all of these settings,
the reduction in tooth wear appears to be linked with the shift from reliance on nondomesticated
to domesticated plants or more intensified use of domesticated plants and changes in associated
food preparation technology. In the Oklahoma–Arkansas and Tennessee regions, the change in
diet involved the adoption of or increased emphasis on maize, a food which was typically
prepared into a soft gruel. Perhaps of equal importance in this setting in relation to the reduction
in tooth wear is the replacement of grinding stone implements used for processing nondomes-
ticated plants with wooden mortars and pestles that were used to process maize (Hinton et al.,
1980; Powell, 1985). The shift from stone to wood in food preparation technology greatly
reduced the harsh abrasive content of foods consumed by late prehistoric Indians, because it
cut down on the number of exogenous materials added to foods.
Details on diet and food preparation in native populations are provided in historical
descriptions for some regions. For example, in Virginia and the Carolinas, Arthur Barlowe
described a late sixteenth century meal as including “some wheate like furmentie, sodden
Venison, and roasted, fish sodden, boyled, and roasted, Melons rawe, and sodden, rootes and
divers kindes and divers fruites” (cited in Hoyme & Bass, 1962:353). Observations such as
these provide important perspectives on the extent to which southeastern North American
tribes consumed soft foods.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
280 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

Old World populations represented by temporal sequences of skeletal remains of earlier


hunter-gatherers and later agriculturalists generally show a change from high wear to
moderate or low wear, regardless of the kind of domesticated plants consumed (various in
Cohen & Crane-Kramer, 2007). For instance, a major dietary transition during the Neolithic in
west-central Portugal involved a shift in subsistence from foraging and fishing to a more
terrestrial-focused diet that included domesticated animals (pig, sheep, goat) and plants
(various grains) (Cunha et al., 2007; Lubell et al., 1994). Quantification of tooth wear reveals
that these populations show a dramatic and rapid reduction in severity of occlusal wear
during this adaptive shift. For Mesolithic individuals from Moita with their lower third molars
newly erupted, 87.5% have lower first molars with occlusal surface enamel worn away
completely, and comparably aged individuals from the Neolithic show significantly less wear
on their lower first molars (Lubell et al., 1994). Similarly, the foraging-to-farming transition
(Mesolithic to Neolithic) in Scandinavia reveals a decline in tooth wear, reflecting the shift
from eating marine foods and some plant carbohydrates to terrestrial foods (meat, some fish,
domesticated grains) (Bennike & Alexandersen, 2007).
Some of the earliest and best documented evidence of agriculture and food production in
post-Pleistocene populations is from the Near East, especially in the record of the foraging-to-
farming transition represented by the earlier Natufian and later Neolithic series spanning
some 5000 years (10 500 BC–5500 BC) in the Levant (Israel). The foragers collected a range of
plants, fished, and hunted. The very latest Neolithic period was the first culture to be fully
focused on domesticated plants (wheat, barley, and legumes) and animals (sheep, cattle, goats,
and pigs). Prior to this time, there was a gradual transition from consumption of fully
nondomesticated plants and animals to the adoption and intensification of farming. In this
setting, there is a clear reduction in severity of occlusal wear, especially in the mandibular
dentition and predominantly in the anterior teeth (Eshed et al., 2006) (Figure 7.7). These
patterns express the overall trend for reduction in wear and the shift in wear dominance from
anterior to posterior teeth.
In a series of sites dating from the period of 12 000–7000 years ago in northern Syria,
plant, animal, and human remains provide information on the complex nature of the shift
from a primarily foraging to an intensive agricultural economy (Molleson & Jones, 1991).
Populations placed a heavy emphasis on at least six domesticated cereals, beginning about
10 000 years ago. Preliminary findings suggest that unlike the settings discussed earlier, the
initial agriculturalists in the early Neolithic express an increase in the severity of occlusal
wear in comparison with earlier foragers (Molleson & Jones, 1991; and see Blau, 2007;
Smith, 1972). This change appears to be related to the shift from less-coarse nondomesti-
cated grains consumed by Mesolithic hunter-gatherers to coarse grains consumed by Neo-
lithic farmers. Both groups used grinding stones to process grains; thus, it is unlikely that
the manner of food preparation led to the increase in wear in the later populations. During
the later Neolithic period, there is a reversal in wear, as most late Neolithic individuals show
reduced wear compared with early Neolithic individuals. This is probably because ceramic
vessels were used in the later period for boiling grains and other foods into soft mushes.
Thus, in this case, innovations in food preparation technology resulted in lessened wear
rates. Analysis of microwear in these populations confirms the reduction in dietary texture
and abrasiveness (Molleson & Jones, 1991; Molleson et al., 1993; and see the following
discussion).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 281

Neolithic
7
Natufian

6
Attrition Level (1–10)

0
I1 I2 C P3 P4 M1 M2 M3 I1 I2 C P3 P4 M1 M2 M3

Mandible Maxilla

Figure 7.7 Changes in dental attrition comparing Natufian (preagricultural) and Neolithic
(agricultural) populations in the southern Levant. The transition to an agricultural
subsistence refocuses attritional wear from anterior to posterior dentition. However, a
general reduction in attritional levels can be observed from Natufian to Neolithic periods.
(Data from Eshed et al., 2006.)

Elsewhere in Asia, limited evidence from a handful of settings shows a general pattern of
decreasing severity of wear with the foraging-to-rice farming transition or rice intensifica-
tion, for example in Japan (Inoue et al., 1986; Kaifu, 1999) and China (Liu et al., 2010;
Pechenkina et al., 2002). Similarly, in Neolithic Thailand, there is a decline in severity of wear
inferred from the prevalence of individuals with advanced occlusal surface wear (Douglas,
2006), and in later contexts where rice cultivation and consumption intensified (Domett &
Tayles, 2007). Throughout East Asia, rice was the dominant domesticated plant, taking on a
central role of diet in later prehistory and the historic period. As with other settings globally,
the general reduction in tooth wear reflects the co-occurrence of two factors, namely the
consistency and texture of the food consumed and the manner in which it is processed for
consumption. With regard to rice and its preparation in Japan, considerable evidence reveals
that rice, when boiled, is soft textured, and other treatments that developed over the course of
time served to reduce the texture to being remarkably soft. Moreover, experimental evidence
of food preparation practices that were reconstructed from ancient Japanese periods and the
food fed to students revealed decreased chewing time and masticatory cycles (Kaifu, 1999).
Moreover, Kaifu (1999) suggests that the adoption of chopsticks first invented in China and
then adopted in Japan sometime between the third and seventh centuries AD resulted in
reduced initial food preparation by the anterior teeth, thus explaining, for this setting, the
reduced anterior tooth wear in agricultural-intensive populations.
A general decline in masticatory demands with agriculture and its intensification has also
been documented via analysis of occlusal wear in northeastern Africa (Greene et al., 1967).
Mesolithic foragers in Nubia have severe occlusal wear, which reduces considerably in the
following agricultural-dependent populations (Meroitic, X-group, Christian group). This

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
282 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

decline in wear had important implications for changing patterns of oral health, as the
reduction of wear coincides with an increase in dental caries and other dental pathological
conditions. Perhaps, then, the continued presence of caries-prone grooves and fissures from
reduced wear in the agricultural groups may have predisposed their teeth to increased decay
with the shift to an emphasis on plant carbohydrates (Greene et al., 1967).
Collectively, these studies underscore the very strong influence of dietary change on the
wear environment, especially in relation to the worldwide transition from food collection to
food production and the later intensification of food production with the focus on domesti-
cated plants. Analysis of dental wear in settings involving other kinds of dietary shifts, within
both hunter-gatherers (Lieverse, Link et al., 2007; Walker, 1978) and agriculturalists (Gua-
landi, 1992), demonstrates appreciable changes in masticatory behavior not involving the
foraging-to-farming transition. Early and late period populations from the Santa Barbara
Channel Islands region saw a dramatic shift in dietary focus – from an emphasis on terrestrial
foods to a reliance on marine foods after AD 1150 (Glassow, 1996; Walker, 1978). In the early
period, terrestrial foods included seeds, prepared with grinding stones, and significant quan-
tities of shellfish (Erlandson, 1994; Glassow, 1996). In the later period, a much greater
emphasis was placed on fishing and hunting of marine mammals (Glassow, 1996; Walker,
1991–1992). The changes in tooth wear in these populations are profound, including a
decrease in wear as measured by dentin exposure area on occlusal surfaces. This trend is
likely related to the change in dietary emphasis and food preparation technology. The use of
milling equipment by earlier populations contributed to relatively greater wear than in the late
prehistoric populations. Similarly, in the Lake Baikal setting of central Asia, there was an
expansion of types of foods eaten in comparison of Kitoi culture (late Mesolithic–early
Neolithic; c. 6800–4900 BC) and later Isakovo-Serovo-Glaskovo culture (middle Neolithic–
early Bronze Age; c. 4200–1000 BC). Analysis reveals consistently high levels of occlusal wear
throughout the temporal sequence (i.e., there was no change in wear over time). The lack of
change is interesting because there had been significant behavioral (increased mobility) and
dietary (increased breadth, but a relatively reduced focus on fish) change in the comparison of
the two periods (Lieverse, Link et al., 2007). On the other hand, individuals associated with
lake shore sites, in both earlier and later periods, had a consistently lower severity of wear,
suggesting circumstances involving that particular context.

Interproximal wear
Bioarchaeologists have given considerably less attention to the study of interproximal wear,
which involves flattened facets at the areas of contact between adjacent teeth. Interproximal
wear results from differential movement of teeth during chewing; the greater the mechanical
forces placed on the teeth during mastication, the greater the amount of interproximal wear
(Hinton, 1982; Wolpoff, 1971). For example, Australian Aborigines eating tough or abrasive
foods have far more interproximal wear than Europeans eating soft, processed foods (Wolpoff,
1971). Hinton (1982) documented a significant reduction in interproximal wear as one moves
in temporal succession from prehistoric Archaic- to Woodland- to Mississippian-period
premolars and molars in the Tennessee River valley (Table 7.2, Table 7.3). This trend suggests
a decrease in masticatory loading of the dentition, a finding consistent with declining
craniofacial robusticity in these populations (Hinton, 1981a). For example, temporomandib-
ular joint size reduces from the earliest to latest periods. This record reveals that although the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 283

Table 7.2 Interproximal wear-facet breadth, PM2/M1, stratified by level of occlusal wear;
prehistoric Tennessee Native Americans (Adapted from Hinton, 1982. All measurements in
millimeters)
Occlusal wear level on lower M1

3 4 5 6 7 8
Period
Archaic 4.9 5.4 5.3 6.1 6.4 7.1
(n) (8) (6) (6) (18) (31) (22)
__
Woodland 4.8 5.3 5.4 5.9 6.0
__
(n) (33) (11) (9) (22) (8)
__ __
Mississippian 4.3 4.6 4.9 4.7
__ __
(n) (43) (22) (10) (6)

Table 7.3 Interproximal wear-facet breadth, M2/M1, stratified by level of occlusal wear;
prehistoric Tennessee Native Americans (Adapted from Hinton, 1982. All measurements in
millimeters)
Occlusal wear level on lower M2

2 3 4 5 6 7 8
Period
Archaic 4.8 5.5 6.0 6.3 6.7 6.7 6.9
(n) (7) (16) (7) (5) (18) (30) (8)
__ __
Woodland 4.2 5.0 5.4 5.7 6.2
__ __
(n) (14) (32) (13) (9) (8)
__ __ __
Mississippian 4.0 4.3 5.1 4.5
__ __ __
(n) (28) (42) (3) (3)

dietary change likely had an important influence on both interproximal wear and craniofacial
robusticity, the more important change may have been related to the manner in which food
was prepared. At the time of contact, native populations were consuming various soft foods,
such as mushes, stews, and puddings (Hudson, 1976; Swanton, 1946). Archaeological and
ethnobotanical evidence of foods and food processing in the earlier periods indicates a greater
coarseness of diet requiring more mechanical force in chewing than in the later periods.
Comparisons of a temporal series spanning the shift from foraging to farming in southern
Ontario likewise show a decrease in interproximal wear (Patterson, 1984). This change is
consistent with Hinton’s model of reduction in masticatory loading in relation to consump-
tion of soft, agricultural-based foods during later prehistory in the Eastern Woodlands.

Occlusal wear patterns


Patterns of occlusal surface wear provide additional perspective on masticatory and dietary
behavior in past human populations. A number of early investigators noted trends in wear
patterns with probable functional significance. In his detailed report on pathological

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
284 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

conditions present in human remains recovered in Nubia, Jones (1910:279, 282) compared
occlusal wear patterns in Predynastic dentitions with later samples (e.g., Christian and Byzan-
tine periods), noting a clear tendency for flat (“level and even”) wear on tooth crowns in the
former, contrasting with “a marked hollowing-out of the centre of the crowns” in the latter.
In order to test the hypothesis that systematic differences in human populations in tooth
wear patterns are related to subsistence and food preparation, Smith (1984) examined flatness
of occlusal wear in hunter-gatherers versus agriculturalists on a worldwide scale. Foragers
included European and Near Eastern Middle and Upper Paleolithic, French Mesolithic, pre-
contact Australian, precontact Eskimo, and Archaic-period Native Americans from Alabama;
agriculturalists from the British and French Neolithic, Nubia, historic Britain (Iron Age,
Anglo-Saxon, Medieval), Mississippian-period Alabama, and late prehistoric American
Southwest Pueblo (Smith, 1984:42). Although the general severity of wear was similar within
the two groups, and foragers overall had relatively more severe wear than agriculturalists;
these differences were not uniform. For example, tooth wear in the Nubian agriculturalists
was as severe as tooth wear in Australian Aborigine hunter-gatherers.
In spite of a great deal of variation in diet and food preparation technology across the
samples, highly consistent differences in the angle of the occlusal wear plane on the man-
dibular first molars were documented. Agriculturalists show higher angles of occlusal surface
wear than hunter-gatherers. For the more advanced wear stages (see Figure 7.8), these
differences approach 10. Overall, hunter-gatherers have flatter wear than agriculturalists.
Moreover, wear on first molars differs by subsistence strategy: wear is cupped in agricultur-
alists and flat in hunter-gatherers (Figure 7.9; and see Eshed et al., 2006; Kasai & Kawmura,
2001; Pinhasi et al., 2008). This compelling analysis shows that wear plane angles and form
reflect the kinds of foods being eaten as well as the manner in which they are prepared. Smith
drew the general conclusion that these differences were related to greater “toughness or
fibrousness” of the diets in foragers than in farmers.
Thus, the pattern of flat and cupped wear that Jones (1910) first observed in his compari-
sons of Predynastic and Christian-era populations in Nubia reflects the distinctions between
the respective hunter-gatherer and agriculturalist groups documented in Smith’s study.
Additional confirmation of Smith’s findings has been provided in other settings undergoing
the shift from food collection to food production. In South Asia, Pastor (1992) compared gross
wear and microwear (discussed later) in a series of prehistoric hunter-gatherers’ lower first and
second molars from the Mesolithic site of Mahadaha in the Ganga River valley and incipient
agriculturalists from the Chalcolithic site of Mehrgarh in the Indus River valley. Diet in the
Mesolithic was comprised of tough, fibrous foods, including nondomestic mammals and wild
grains, roots, and other plants that were processed in stone querns. Chalcolithic populations
consumed domestic animals (cattle, sheep, goats) and several varieties of wheat and barley. In
addition to showing a marked decline in wear severity, the pattern of wear is strikingly similar
to Smith’s hunter-gatherer and agriculturalist samples for the Mesolithic and Chalcolithic,
respectively: Mesolithic teeth are worn flat at slight angles, and Chalcolithic teeth are worn at
a greater angle with distinctive cupping.
In the previously mentioned Portuguese populations, molar occlusal surfaces also show a
temporal change from flat to cupped wear as one moves from the Mesolithic to Neolithic sites
(Lubell et al., 1994). This study also serves to illustrate the potential variation of wear patterns
in human populations as wear angles in Mesolithic foragers are much greater than in later

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 285

(a)

(b)

Figure 7.8 Lateral views of tooth wear in Nubian A-group agriculturalist (a) and Eskimo
hunter-gatherer (b), showing greater angle of wear plane in the former than the latter.
(From Smith, 1984; reproduced with permission of author and John Wiley & Sons, Inc.)
(A black and white version of these figures will appear in some formats. For the color
version, please refer to the plate section.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
286 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

(a)

(b)

Figure 7.9 Occlusal view of tooth wear in Nubian X-group agriculturalist (a) and Mesolithic
hunter-gatherer (b) showing cupped occlusal wear on first molars in the former but not the
latter. (From Smith, 1984; reproduced with permission of author and John Wiley & Sons,
Inc.) (A black and white version of these figures will appear in some formats. For the color
version, please refer to the plate section.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 287

Neolithic farmers. The reasons for the differences from Smith’s findings are largely unknown.
Lubell and coworkers (1994) speculate that differences between the Portuguese populations
and the samples studied by Smith, especially in age composition and molar crown morph-
ology, may be important considerations. Thus, while a common pattern includes flat wear in
hunter-gatherers and oblique wear in agriculturalists, variability is present in some groups
(Schmucker, 1985).
Other types of wear demonstrate significant variation in masticatory behavior in past
human populations. Analysis of prehistoric foragers and farmers from a wide range of
contexts reveals important differences in forms and severity of anterior tooth wear compared
to posterior tooth wear. In general, in hunter-gatherers, severity of anterior tooth wear is
greater than, or equal to, severity of wear on posterior teeth. In farmers, anterior wear is
appreciably less severe than posterior wear (Bennike & Alexandersen, 2007; Brown & Molnar,
1990; Dahlberg, 1963; Deter 2009; Duray, 1996a, 1996b; Eshed et al., 2006; Hinton, 1981d,
1982; Kaifu, 1999; Molnar, 1971, 1972; Schmucker, 1985; Sciulli, 1997; Tomenchuk &
Mayhall, 1979; Turner & Cadien, 1969; van Reenen, 1992; Walker & Erlandson, 1986). The
form of wear on the anterior teeth is distinctive: prehistoric foragers show a characteristic
rounded wear, whereas the agriculturalists have cupped wear (Hinton, 1981d). The consist-
ency of these findings indicates that tooth wear is mediated by masticatory behaviors. The
relatively heavier use of the anterior teeth in hunter-gatherers is in line with ethnographic
reports of tooth use in Australians and Eskimos, in both dietary and non-dietary functions.
The distinctive labially oriented rounded wear in hunter-gatherer dentitions is likely a
function of the use of the incisors and canines in various extramasticatory activities (e.g.,
hide preparation; and see following discussion). The distinctive cupping wear on agricultur-
alist incisors and canines appears to be especially prominent in individuals who have lost
posterior teeth prior to death. Thus, this anterior wear pattern may represent an alteration of
tooth use arising from the loss of posterior teeth, namely from the use of the incisors and
canines for a combination of food preparation and mastication.

Social differences in pattern and severity of wear: sex and status


Sex differences in pattern and severity of wear indicate behavioral variability in tooth use for
a range of human populations. Adult female incisors and canines are more worn than male
incisors and canines in the Mesolithic population from Skateholm, Sweden (Frayer, 1988b).
Similar differences are observed in native populations from North America (Clement &
Hillson, 2012; Reinhard et al., 1994), South America (Da-Gloria & Larsen, 2014), and South
Africa (Morris, 1992). In these settings, differences in anterior tooth wear suggest gender-
specific behaviors. For example, historic evidence indicates that Omaha women from the
Great Plains were responsible for hide processing, and probably used their front teeth in the
activity. Among Inuit in northern North America, clear sex differences in use of the anterior
dentition – with women using the incisors and canines to soften animal hides – is well known
ethnographically. In South Africa, some San foragers use their anterior teeth to prepare plant
fibers to make ropes (van Reenen, 1964). The study of archaeological dentitions from this
setting suggests that these activities were primarily female responsibilities.
In Australian Aborigines, differences in female and male occlusal surface wear in a number
of settings appear to be associated with gender role-based masticatory and extramasticatory
activities (Littleton et al., 2013; Richards, 1984). Dental wear in South Australian foragers

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
288 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

shows differences between the sexes, suggesting especially gendered patterns of wear. For
example, Euston females have a flat wear plane on the anterior teeth and maxillary premolars
(Littleton et al., 2013). Similarly, Gillman females lack heavy anterior and premolar wear, but
some males do. In contrast, in Kaurna foragers, females, but not males, expressed heavy wear
on incisors and canines (Littleton et al., 2013; Richards, 1984). These findings strongly suggest
gendered roles as they pertain to extramasticatory tooth use. Although the reasons for this
variation are unknown, they attest to the presence of highly variable wear patterns between
closely related populations in what is often characterized as a homogeneous region (Littleton
et al., 2013; Molnar et al., 1989).
Analyses of tooth wear patterning in other settings generally show more wear on occlusal
surfaces in males than females (Da-Gloria & Larsen, 2014; Douglas, 2006; Esclassan et al.,
2008; Fujita & Ogura, 2009; Kaifu, 1998, 1999; Meng et al., 2011; Pechenkina et al., 2002).
Regardless of the differences – male wear greater than female wear, or female wear greater
than male wear – the record shows evidence for clear sex differences in the wear environment,
suggesting differences in either diet or use of teeth, or both, are a part of gender identity as it
relates to masticatory and extramasticatory function.
Gross wear differences by social group and rank have been identified in past societies. In
the Medieval Edo period of Japan, members of the elite Shogun class had virtually no occlusal
surface wear, unlike lower status, non-Shogun individuals (Suzuki, 1969). This indicates a less
mechanically demanding, less abrasive diet in the elite than in the remainder of Japanese
society. The presence of narrow faces, reduced size of the maxillae and mandibles, and gracile
masticatory muscle attachment sites in Shogun individuals corroborates a reconstruction of
the consumption of soft, processed foods by the elite.

7.4.2 Extramasticatory wear


Some of the greatest mechanical demands placed on the dentition involve the use of teeth as
“tools” in nonmasticatory functions. Milner and Larsen (1991:364) note that “teeth can show
the effects of a wide variety of activities unrelated to eating that result in unusual, and at
times highly distinctive, patterns of abrasion, crown fractures, or traumatic tooth loss.” Until
recently, a variety of human populations have used their teeth in extramasticatory ways, and
in no other group is this expressed as well as among circumpolar populations, such as the
Inuit (Cybulski, 1974; Leigh, 1925; Merbs, 1983; Molnar, 1972; Pedersen, 1952; and many
others). The following discussion considers unintentional changes in teeth arising from
extramasticatory activities.
Modifications involving transversely oriented grooves on worn occlusal surfaces of per-
manent mandibular incisors and canines are especially illustrative of the role of teeth as tools
in a number of foraging human groups in prehistoric North America. Five prehistoric older
adult males from the Great Basin in west-central Nevada display well-defined single or
multiple grooves located on anterior teeth (Larsen, 1985) (Figure 7.10). The grooves are
generally polished and rounded; scanning electron microscopy (SEM) analysis reveals a series
of fine scratches lying parallel to each groove’s main axis (Larsen, 1985). Similar types of
grooves are present on occlusal surfaces of prehistoric forager dentitions from California
(Schulz, 1977), Prince Rupert Harbour in British Columbia (Cybulski, 1974), and central Texas
(Bement, 1994). The high degree of polish and the orientation of grooves in these samples

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 289

Figure 7.10 Adult mandibular dentition showing occlusal surface grooves; Great Basin,
Nevada. (Reproduced with permission of John Wiley & Sons, Inc.) (A black and white
version of this figure will appear in some formats. For the color version, please refer to
the plate section.)

indicate that, as in the Great Basin setting, some type of flexible material had been passed
transversely over the anterior teeth in a repetitive and habitual fashion, especially in process-
ing materials such as sinews for bow strings or plant fibers for cordage or basketry.
Ethnographic and historical evidence provides corroborative support for possible uses of the
dentition that may have resulted in grooves. In the southwestern margin of the American Great
Basin (Death Valley), Coville (1892) described native Panamint women preparing sumac and
willow for materials to be used in making “wicker-work utensils”: A woman “selects a fresh
shoot. . .bites one end so that it starts to split into three nearly equal parts. Holding one of these

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
290 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

parts in her teeth and one in either hand, she pulls them apart, guiding the split with her
fingers. . .Taking one of these, by a similar process she splits off the pith. . .leaving a pliant, strong,
flat strip of young willow or sumac wood” (Coville, 1892:358). Greenland Eskimo women pull thin
cords of animal tendon across the clenched anterior teeth in order to moisten and soften the sinew
(Pedersen and Jakobsen, 1989). Exclusively older women (>40 years of age) are responsible for
the preparation of sinew in this manner. Interestingly, in archaeological dentitions, only older
adult female dentitions display these “sinew grooves” (Pedersen & Jakobsen, 1989). The presence
of grooves in females only in Eskimos and in males only in Great Basin Native Americans suggests
that activities causing these grooves were gender-based (but see Schulz, 1977).
While a range of settings show the presence of occlusal surface grooves in anterior teeth in
New World foragers, these dental modifications are increasingly becoming recognized in Old
World foragers and farmers. Early Holocene foragers from north Africa (Uan Muhuggiag,
Libya: Minozzi et al., 2003), pre-Neolithic (Natufian) incipient farmers from Israel (Bocquentin,
2007; Eshed et al., 2006) and Jordan (Webb & Edwards, 2002), Neolithic farmers from Syria
(Abu Hureyra: Molleson, 1994) and Poland (Lorkiewicz, 2011), and Byzantine dentitions from
western Asia (Kovuklukaya, Turkey: Erdal, 2007) also display distinctive patterns of occlusal
surface grooving. In these settings, grooves may have been produced by dragging plant fibers
across the surface of the teeth as part of the process of preparing materials for basket production
(Eshed et al., 2006; Lorkiewicz, 2011; Minozzi et al., 2003; Molleson, 1994) or yarn for making
clothing (Erdal, 2007). Preparation of tough plant materials for utilitarian purposes also may
explain the presence of occlusal surface grooves in prehistoric manioc agriculturalists in
St. Thomas, US Virgin Islands (Larsen et al., 1998; Larsen, Teaford et al., 2002).
Alterations of anterior teeth involving notching of the mesial or distal occlusal margins
have been observed in southeastern US samples from Tennessee (Blakely & Beck, 1984) and
the Georgia coast (Larsen & Thomas, 1982). In the latter case, an individual shows wear on the
mesial corner of a maxillary right first incisor that resulted from extramasticatory use, such as
from clamping fishing nets or from processing plant materials. The presence of cumulative
wear indicates that the teeth were not intentionally altered for personal ornamentation, such
as with purposeful drilling.
Lingual surface wear on maxillary incisors and canines is another highly distinctive pattern
of tooth wear involving extramasticatory behaviors (Figure 7.11). Most (39/46) adult crania
from the preceramic (4200–3000 yBP) site of Corondó, Rio de Janeiro State, Brazil, display
pronounced lingually oriented flat wear (Turner and Machado, 1983). There is no corresponding
wear on the lower teeth. Called lingual surface attrition of the maxillary anterior teeth
(LSAMAT), the orientation of wear striations strongly suggests that some kind of extramasti-
catory use of the teeth caused the wear pattern. Its presence in adult females and males suggests
that it is not associated with a gender-specific activity. The presence of moderate lingual
wear in anterior teeth of older children (beginning at about 10–11 years of age) indicates the
time in life when the behaviors that cause the wear commence. Given the high prevalence of
caries in this group (10.7% of teeth) and the known heavy reliance on manioc – a cariogenic
carbohydrate requiring extensive preparation prior to consumption – the teeth appear to have
been used to process plants for consumption (Turner & Machado, 1983). Support for this
argument has been provided by the study of additional prehistoric dentitions from other New
World settings, including Panama (Irish & Turner, 1987), Belize (Saul & Saul, 1991), Texas
(Hartnady & Rose, 1991), and the US Virgin Islands (Larsen et al., 1998; Larsen, Teaford et al.,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 291

Figure 7.11 Lingual wear on maxillary incisors; Çatalhöyük, Turkey. (Photograph by


Scott Haddow.) (A black and white version of this figure will appear in some formats.
For the color version, please refer to the plate section.)

2002). In these settings, the form of wear, reliance on manioc (or other fibrous plants), and
high prevalence of dental caries argue that these modifications are due to the habitual use of the
upper teeth in preparation of abrasive plant material. Because these studies indicate that wear
occurs mostly on the upper teeth, the material being manipulated must have been placed between
the upper teeth and tongue and drawn across the upper teeth in a back-to-front motion. With
regard to the Brazilian populations, Turner and Machado (1983) suggest that the upper teeth were
used to shred or peel tuberous plants (e.g., manioc) “comparable to the modern way we eat
artichokes – by pulling and planing the edible petals across the occlusal surfaces of our anterior
teeth” (1983:128; and see Irish & Turner, 1987). Microwear analysis of incisors from the Tutu
population from the Virgin Islands shows that minute striations caused by the passing of material
across the tooth surface involved a back-to-front motion (Larsen et al., 1998), thus adding
additional confirmation of Turner and Machado’s artichoke hypothesis.
A similar pattern of lingual wear in maxillary anterior teeth has been identified in a range
of Old World settings, including eastern Europe (Lorkiewicz, 2011), the Mediterranean basin
(Harper & Fox, 2008), Africa (Irish & Turner, 1997), western Asia (Lukacs & Pastor, 1988), and
northern China (Liu et al., 2010; Meng et al., 2011; Miao et al., 2013; Pechenkina et al., 2002).
Each of these settings has very different dietary and subsistence technology, suggesting
alternative uses of the lingual surfaces of teeth in extramasticatory functions. For example,
Lukacs and Pastor (1988) suggest that lingual wear in a Neolithic dentition from Mehrgarh,
Baluchistan, resulted from the use of teeth for the processing of animal skins. However, the
strong association between caries and lingual wear in at least one setting in historic-period
West Africa suggests a link between high carbohydrate diet (manioc) and dental processing
(Irish & Turner, 1997). On the other hand, the significant presence in low-carbohydrate

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
292 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

contexts reveals the varied associations of this behavioral marker (Miao et al., 2013; Pechen-
kina et al., 2002). It is likely that this form of wear relates to preparation of fibers generally,
for either dietary purposes (the dominant pattern in the New World) or material culture.

7.4.3 Microwear
Microwear, or microscopic damage on teeth by materials processed in the oral cavity, provides
essential insight into dietary adaptation and masticatory function. The development of a number
of tools for accessing intricacies of dietary adaptation, tooth use, and masticatory behavior make
possible more effective ways of understanding use wear (Teaford, 1991; Ungar et al., 2008). The
depth of focus and highly detailed resolution in SEM analysis (Teaford, 1991), and more recently,
microwear texture analysis (Ungar et al., 2008) have expanded our understanding of diet and
tooth use for a wide range of primates, including humans (Figure 7.12). Because the analysis of
microwear features (e.g., pits and scratches) requires high magnification (e.g.,  500), only small
areas can be assessed at any one time. Texture analysis, on the other hand, provides a more

(a)

Figure 7.12a Scanning electron micrographs of maxillary permanent first molar


occlusal surfaces ( 500). Top: deep pits and wide grooves in the forager (Marys Mound,
St. Catherines Island, Georgia). Bottom: slight pits and narrow grooves in the farmer
(Santa Catalina de Guale de Santa Maria, Amelia Island, Florida). (From Teaford, 1991.
Reproduced with permission of WILEY-LISS.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 293

(b)

Figure 7.12b Scanning electron micrograph ( 100) of section of occlusal surface of


second mandibular molar from an Archaic-period forager from the Meyer site, Indiana.
(Image courtesy of Melissa Zolnierz and Christopher Schmidt.)

(c)

Figure 7.12c Confocal microscope 3D image ( 100) of same location on same tooth. (Image
courtesy of Melissa Zolnierz and Christopher Schmidt.) (A black and white version of this
figure will appear in some formats. For the color version, please refer to the plate section.)

readily interpretable three-dimensional image of occlusal surface microwear, along with a


scaling system that makes comparisons more straightforward and with less error than in
traditional SEM analysis. Moreover, unlike the traditional microwear analysis, texture analysis
provides significantly larger samples and more repeatable results (Ungar et al., 2008).
Studies of extant and fossil primates and hominins show that qualitative and quantitative
analysis of occlusal surface microwear features provides useful information on dietary
behavioral differences among and between animal species along with a range of implications
for diet and environment (Crompton et al., 1998; Daegling & Grine, 1999; Gordon, 1984;
Grine & Kay, 1988; Kay, 1987; Ryan, 1981; Ryan & Johanson, 1989; Scott et al., 2012;
Silcox & Teaford, 2002; Teaford & Walker, 1984; Ungar, 1994; Walker, 1980, 1981). In all of
these studies, there are several potential drawbacks in SEM analysis, whether determinations
of wear are made in comparison of different species or between different time periods within

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
294 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

species. The key limitation of microwear analysis is that it provides a record of tooth use that
documents masticatory or dietary behavior for the time immediately prior to death, from
hours to a few days, otherwise known as the “Last Supper” phenomenon (Grine, 1986).
However, the generally high degree of consistency in various analyses indicates the import-
ance of this kind of study in assessing patterns of tooth use and in drawing inferences about
masticatory behavior (Teaford, 1991; Ungar et al., 2008).
Experimental research indicates that teeth of animals and humans fed soft or nonabrasive diets
have fewer microwear features than animals and humans fed hard or abrasive diets (Teaford,
1991; Teaford & Lytle, 1996). Like macrowear, these differences are influenced by the nature of
the food itself and by extraneous elements, such as soil particles, which may be incorporated in
the food prior to its consumption. Microwear variation in archaeologically recovered humans
indicates differences related to food texture and diet within specific groups (Danielson &
Reinhard, 1998; Fine & Craig, 1981; Ma & Teaford, 2010; Mahoney, 2007; Marks et al., 1988;
Organ et al., 2005; Puech et al., 1983) or in relation to subsistence change (Bullington, 1991; Hojo,
1989; Homes Hogue & Melsheimer, 2008; Mahoney, 2005, 2006, 2007; Rose, Anton et al., 1991;
Schmidt, 2001; Teaford, 1991; Teaford et al., 2001; Ungar & Spencer, 1999).
Occlusal microwear on molars from Early Mississippian-period pre-maize populations and
Middle Mississippian-period (Zebree site: AD 900–1200) maize-dependent populations in the
central Mississippi River valley underwent a change from the presence of numerous large
striations to few large striations; occlusal surfaces go from being rough to smooth (Rose,
Anton et al., 1991). Rose and coworkers (1991:17) interpret the striking change in surface
topography as indicating “a radical reduction in abrasive particles associated with changes in
both food content and preparation.” They note increased prevalence of dental caries, from 2.6
to 3.5 lesions per dentition, and less negative δ13C values, indicating that the microwear
alterations resulted from a change in food consumption.
Similarly, an association between dietary change and microwear has been documented in a
temporal series from the southeastern US Atlantic coast and interior northern Florida,
including precontact hunter-gatherers, later precontact farmers, and early and late mission-
era populations living in Catholic missions (Teaford, 1991; Teaford et al., 2001). The pre-
contact populations underwent a transition from foraging to maize farming in late prehistory,
and the later contact populations were largely intensive maize agriculturalists. Overall,
precontact molars display narrower scratches on the occlusal surfaces than do the mission
molars. These changes in microwear reflect a reduction in abrasiveness of diet during the
contact period, suggesting that maize-based diets were of a softer-textured nature than the
foraging-based diets. On the other hand, the analysis revealed considerable variation in
microwear features in the region, reflecting the highly localized aspects of diet and habitat,
patterns which reflect similar conclusions drawn in the analysis of carbon and nitrogen stable
isotopes (Hutchinson et al., 1998; Larsen, Griffin et al., 2001). At San Luis de Apalachee (AD
1656–1704), an economically important center at the western end of the mission system in
Spanish Florida, molars express an unusually high volume of surface pits and scratching
(Organ et al., 2005). Historical sources point to the importance of cattle in this setting and of
maize farming. However, the low prevalence of dental caries and preliminary isotopic
evidence for lower meat consumption than at other missions suggests that the unusual
microwear reflects a diet relatively higher in meat. If meat played an important role in diet,
then it would have been relatively tough, perhaps with soil and/or grit inclusions.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 295

The number of microwear features on occlusal surfaces of deciduous incisors and molars
from Middle Woodland (50 BC–AD 250) incipient pre-maize agriculturalists and Mississippian
(AD 1000–1350) maize agriculturalists in the lower Illinois River valley of western Illinois
increased with age during both periods (Bullington, 1991). Reflecting the difference in use of
anterior and posterior teeth, the incisors have relatively more scratches than pits and the
molars have more pits than scratches. These differences indicate the use of the incisors in
initial processing and the molars in crushing hard objects. The earlier group focused on wild
plants and animals in combination with a reliance on various starchy seeds and those with
hard seed coats. The later group continued to utilize these food items, but with their partial
replacement by maize. In addition, these foods were likely boiled in ceramic vessels for long
periods of time, thus reducing the toughness of foods consumed by prehistoric populations
(Bullington, 1991). Unlike the changes observed in the Mississippi River valley and the
Georgia Bight, comparisons of the Middle Woodland and Mississippian groups revealed no
differences in microwear. Bullington (1991) suggests that the lack of temporal change may
reflect the fact that her study focuses exclusively on young juveniles; other investigations
focus on adult microwear. There is a strong similarity in microwear between the foragers and
farmers overall, but the two groups are distinctive in the youngest cohort (c. 6–12 months).
Mississippian teeth have lower feature frequencies than Middle Woodland teeth. This suggests
that very young children in the later period were consuming softer foods than very young
children in the earlier period.
Similar changes in microwear in relation to the shift from hard- to soft-textured diets have
been identified in a number of settings in the Old World. Comparison of canines and molars
from two Neolithic (c. 2000 BC) and two historic (AD 1750–1800) dentitions from western
coastal Kyushu, Japan shows a change from generally long, wide striations to thin, narrow
striations (Hojo, 1989). Hojo (1989) suggests that the later diet included smaller abrasives than
did the earlier diet.
Change in occlusal microwear in relation to the shift from foraging to dependence on cereal
grains in western Asia has been investigated via comparison of dentitions from the previously
discussed Neolithic settlement at Abu Hureyra in the Euphrates River valley in northern Syria
(Molleson, 1994; Molleson & Jones, 1991; Molleson et al., 1993). Microscopic analysis of
molars provides corroborative evidence for the study of gross wear. Comparison of pit
diameter, feature frequency, pit density (percentage determined from number and size of
pits), and ratio of striations and other linear features to pits shows a clear shift in microwear in
deciduous and permanent teeth through time. The change in wear is especially conspicuous in
the shift from foraging in the Mesolithic to early agriculture in the pre-pottery Neolithic,
being accompanied by a dramatic increase in microwear feature density. Molleson and
coworkers (1993) attribute this trend to a shift from a relatively soft diet, based on the
consumption of roots and small, wild grains, to a few cereal types prepared with grinding
stone equipment. In the later Neolithic, there is a reduction in feature density and pit diameter,
reflecting a return to softer diets, namely involving cooking of food with ceramic vessels and
the increased consumption of meat. Clearly, the adoption of pottery and its use in cooking
food to a soft consistency had great implications for the manner in which teeth were worn.
The trends identified in Syria are broadly similar to the temporal trends in microwear in the
southern Levant. In this setting, analysis of a small sample of Upper Paleolithic individuals
from Ohalo II (23 500–22 500 yBP) reveals the presence and high frequency of long, narrow

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
296 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

scratches characteristic of populations eating tough, abrasive diets (Mahoney, 2007). Maho-
ney (2007) speculates that this represents consumption of items requiring shearing, such as
wild plants. The infrequency of pits suggests a lack of a hard, brittle diet that would have
required strong compressive forces in the jaws and teeth. In the late Pleistocene and early
Holocene Natufian foragers, however, the molars have various distinctive large pits and wide
scratches, reflecting consumption of hard foods requiring compressive forces in chewing.
Various factors likely explain this temporal change, but the use of grinding stones and the
transfer of quartz particles to the food is the likely cause of this change from earlier popula-
tions (Mahoney, 2005, 2006, 2007). In the early Neolithic populations, the pits became smaller
and the scratches narrower than in the preceding Natufian populations, reflecting a decline in
texture and toughness of foods consumed, similar in many respects to the Upper Paleolithic
individuals from Ohalo II. However, in the later Neolithic period, the pits and scratches
became distinctively larger, a pattern suggesting an increase in toughness of foods consumed,
as in the earlier Natufian population. This pattern indicates an increase in hardness of diet in
the Levantine Neolithic. Overall, there were trends in dietary adaptation, not representing an
increase or decrease in toughness and abrasiveness, but rather, a mosaic of modifications
reflecting both types of foods consumed and how they were prepared for consumption.
A parallel development for a shift from foraging to farming has been documented via
microwear analysis in South Asia, based on the analysis of dentitions from the Mesolithic
(Mahadaha site; 8000–1000 BC), the pre-pottery Neolithic (Mehrgarh site; 7000–6000 BC),
incipient agricultural Chalcolithic (Mehrgarh site; 4500 BC), and urban agriculturalist Har-
appa culture (Harappa site; 2500–2000 BC) (Pastor, 1992, 1993). Comparative analysis of
microwear in permanent first and second molars indicates a generally similar pattern of
increase in features (pits, scratches) as Abu Hureyra. In general, trends in microwear indicate
the higher prevalence of finer scratches in the hunter-gatherers and incipient agriculturalists,
whereas microwear features become coarser in more intensive agriculturalists. The latter
characteristic reflects an increased consumption of cereal grains and use of grinding equip-
ment in the Neolithic and Chalcolithic. The use of stone grinding equipment and its influence
on microwear is illustrated by comparisons of microwear in a living American fed a typical
soft diet with the effects of sandstone-ground maize (Teaford & Lytle, 1996). These compari-
sons reveal that microwear on a maxillary molar increased by 30 times in the consumption of
ground maize over a period of a week. This study supports the observation that food
preparation technology has a profound effect on tooth wear.
Another implication for the research on economic transitions from microwear analysis and
other approaches to functional analysis of the craniofacial complex is the trend for reduction
in hardness and texture of diet, albeit with some variation (Mahoney, 2006, 2007). Presum-
ably, this tendency would be especially pronounced in those members of populations in
higher socioeconomic groups. In late eighteenth- and early nineteenth-century Baltimore,
MD, comparison of occlusal microwear from the Christ Church cemetery and Potter’s Field
East provides an opportunity to characterize and interpret patterns of tooth use in wealthy
and poor persons living in the same habitat, but with different diets (Ma & Teaford, 2010).
Individuals interred in both cemeteries would have consumed seafood, pork, and maize,
among other foods, but the foods consumed by individuals associated with higher-status
Christ Church would have been of a higher quality, prepared in sauces and wine, and they
would have had access to various fruits and vegetables, certainly more frequently than those

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 297

individuals associated with Potter’s Field East (Carson, 1985; Steckel, 1999). Analysis of
microwear, however, reveals no clear differences in how this culinary variation would have
impacted tooth surfaces. That is, analysis of microwear features – pits and scratches – reveals
that while diets were free of abrasives, there are no discernible differences between the two
socioeconomic groups. It may be that traditional microwear analysis based on scanning
electron microscopy is not robust enough to provide the kind of detail on diet and tooth use
needed to discern subtle differences in foodways between the two groups. Importantly,
however, the approach clearly reveals the nature of diets in the nineteenth-century urban
setting of the mid-Atlantic: rich and poor alike enjoyed an abrasive-free diet.
Virtually all of the aforementioned microwear analyses focus on the occlusal surfaces of
teeth. Buccal surfaces have the potential to provide important perspective on tooth use because
wear on them is not influenced by contact between opposing teeth (Ungar & Teaford, 1996). The
importance of buccal surfaces is emphasized in research by Lalueza Fox and coworkers (1996)
on a wide range of past and recent human populations, including hunter-gatherers, pastoral-
ists, agriculturalists, and fossil hominins. Analysis of buccal surfaces of premolars and molars
in these samples reveals that foragers with high-meat diets have fewer and more vertically
oriented striations than agriculturalists. Greater frequency of buccal striations in plant-
consumers may be due to the presence of abrasive phytoliths (Lalueza Fox et al., 1996).

7.4.4 Tooth damage due to masticatory and extramasticatory use


Dentitions from a range of archaeological contexts display damaged enamel, ranging from
slight chipping of occlusal margins to fractures and other related conditions (Bonfiglioli et al.,
2004; Hansen et al., 1991; Lukacs, 2006; Molnar, 2008; Schwartz et al., 1995; Scott & Winn,
2011; Turner & Cadien, 1969; reviewed in Milner & Larsen, 1991). Missing enamel from molars,
ranging from loss of tiny pieces to exfoliation of large blocks, is the most common type of
damage (Figure 7.13). Few systematic investigations of gross premortem damage to teeth, with
regard to either prevalence or distribution of damaged teeth, have been undertaken. This is
especially surprising given the value of studies of occlusal surface damage observed micro-
scopically for masticatory behavioral inference (see earlier discussion), as well as some of the
well-known patterns of crown fracture seen in populations with excessive masticatory
demands (e.g., Eskimos: Hansen et al., 1991; Pedersen, 1947; Schwartz et al., 1995; Scott &
Winn, 2011). An important exception is the comparison of prehistoric and protohistoric Aleuts,
Eskimos, and northern Native Americans, which demonstrates clear distinctions in patterns of
tooth use (Turner & Cadien, 1969). Turner and Cadien (1969) observed a pattern of dental
damage reminiscent of modifications observed on the edges of chipped stone tools. The damage
appears to be due to use of teeth in heavy masticatory functions (e.g., crushing of bone) and in
demanding extramasticatory activities (e.g., hide preparation). Inuit have an unusually high
frequency of chipping damage in teeth (79%) in comparison with Aleuts (22.8%) and Native
Americans (18.4%). Among other Inuit groups, variation in prevalence of chipping reflects
different types of foods consumed and the manner of their preparation. Illustrating this
variation in food and subsistence technology, prevalence of macrodamage varies widely,
ranging from low values in Kodiak Islanders consuming predominantly fish (37.9%) to high
values in Sadlermiut populations consuming caribou and sea mammals (87.5%). Much of this
damage is molar dominant, in part owing to the fact that processing with teeth in these cold

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
298 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

Figure 7.13 Chipped mandibular incisors from the Cathedral of Santa Maria in Vitoria,
Spain, c. AD 1400–1500. (Photograph courtesy of G. Richard Scott.) (A black and white
version of this figure will appear in some formats. For the color version, please refer to the
plate section.)

settings is associated with oral processing and consuming frozen or otherwise tough foods
requiring substantial reduction. Unlike these groups expressing chipping mostly in the poster-
ior dentition, populations from Medieval Norway and Medieval and post-Medieval Spain
express damage but mostly in the anterior dentition (Scott & Winn, 2011) (Figure 7.14). These
differences highlight significant and important variation in how these modern populations
process food and in use of the teeth as an important tool.
As with the Inuit, damaged teeth are also relatively commonplace in other settings. For
example, at Taforalt, Morocco, chipping of teeth is present in 94% (31/33) of individuals, both
in anterior teeth (21.1%) and posterior teeth (32.8%) (Bonfiglioli et al., 2004; and compare
with Milner & Larsen, 1991). While the chipping is likely related to the use of teeth for
processing tough, perhaps uncooked foods, these terminal Paleolithic foragers also practiced
tooth avulsion involving purposeful extraction of anterior teeth, especially permanent inci-
sors. The consequent reliance on the posterior teeth for processing normally involving the
anterior teeth may have contributed to an elevated prevalence of damaged molars (Bonfiglioli
et al., 2004). The considerably greater prevalence of dental chipping in males than in females
in this and most other settings where the condition has been documented suggests a division
of labor and/or difference in diet between sexes (Scott & Winn, 2011). Similarly, prehistoric
Hawaiians display strong differences in dental chipping, but with greater frequency for
women than men (18% vs. 6%). In this setting, the higher prevalence in women is consistent
with ethnographic accounts of women habitually using their teeth for various extramastica-
tory tasks, such as cord production, leaf cutting, and nut cracking. In these populations,
maxillary and mandibular tori are also more prevalent in females than males, thus supporting
a functional interpretation of these skeletal features.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
7.4 Dental wear and function 299

100 Prehistoric St. Lawrence Island Inuit Maxilla

90 Mandible

80

70
Chipped teeth (%)

60

50

40

30

20

10

0
LM3 LM2 LM1 LP4 LP3 LC LI2 LI1 RI1 RI2 RC RP3 RP4 RM1 RM2 RM3

100 Norway (11th–16th Century) Maxilla


Mandible
90

80
70
Chipped teeth (%)

60
50
40
30
20
10

0
LM3 LM2 LM1 LP4 LP3 LC LI2 LI1 RI1 RI2 RC RP3 RP4 RM1 RM2 RM3

100 Spain (11th–19th Century) Maxilla


Mandible
90

80
70
Chipped teeth (%)

60
50
40
30
20
10

0
LM3 LM2 LM1 LP4 LP3 LC LI2 LI1 RI1 RI2 RC RP3 RP4 RM1 RM2 RM3

Figure 7.14 Differences in mandibular and maxillary tooth chipping. Prehistoric


St. Lawrence Island Inuit populations show consistently greater chipping of posterior
teeth (molars), whereas Norwegian and Spanish populations exhibit higher percentages
of chipped anterior teeth (incisors). These differences reflect variability in the level and
types of food processing and tooth tool use. (Data from Scott & Winn, 2011.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
300 Masticatory and nonmasticatory functions: craniofacial adaptation to mechanical loading

Few researchers have documented temporal change in macrodamage to teeth, especially


with respect to major adaptive shifts or change in masticatory behavior (see discussion in
Milner & Larsen, 1991). Comparison of trauma (chipping and fracturing of tooth crowns) in
Middle Woodland hunter-gatherers (LeVesconte Mound site; c. AD 230) and Late Woodland
and historic Iroquois maize agriculturalists (Bennett site [AD 1270]; Kleinburg Ossuary [AD
1600]) reveals a substantial decline in trauma, from 42.9% (LeVesconte Mound) to only 7.4%
(Bennett site) of permanent teeth (Patterson, 1984). This reduction in dental trauma is
consistent with decreases in macrowear reported by Patterson (1984), relating to the dietary
shift to eating softer-textured, maize-based foods. These findings strongly suggest that a
reduction in abrasives in food and food consistency in general lead to less chipping and
fracturing of teeth.

7.5 Summary and conclusions


A great deal of craniofacial and dental variation can be linked directly to the mechanical
environment and the role that diet and subsistence technology play in shaping it. Cranio-
facial form is strongly influenced by functional demands, especially in the lower face and
vault. In the Holocene, the change to a more globular-shaped cranial vault in many areas of
the world appears to represent a compensatory response to the decrease in functional demand
as foods became softer. This conclusion is underscored by comparisons of archaeological
populations that consumed hard-textured foods with populations that consumed relatively
soft-textured foods. Food texture is influenced by the nature of the food itself or the manner
in which it is processed, or both. A range of other data drawn from studies of age changes in
craniofacial size and morphology, occlusal abnormalities, tooth size, gross wear arising from
masticatory and nonmasticatory functions, microwear, and dental trauma provide compelling
evidence for the important role of the dentition and craniofacial skeleton in masticatory and
nonmasticatory behaviors. More importantly, this record provides crucial insight into human
biocultural adaptation and the interface between the masticatory apparatus and craniofacial
variation in recent humans.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 11:31:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.009
Isotopic and elemental signatures of diet,
8 nutrition, and life history

8.1 Introduction
Documentation of past foodways is not only intrinsically valuable, but it also provides the
requisite resource context for evaluating effects of nutrition on growth and development,
assessing stress and disease from paleopathological indicators, and interpreting skeletal
adaptation, among other topics discussed in the foregoing chapters. There is a range of
conventional approaches for characterizing past diets and inferring nutritional outcomes,
including analysis of plant and animal remains, coprolites, pottery residues, and tools used for
food production and consumption. These approaches provide a record of what foods were
consumed, but do not give a representation of the proportions of foods or food classes
consumed by past populations. The poor preservation of plants in many archaeological
contexts, for example, inhibits documentation of their role in diets in many settings. Food
refuse is often subject to preservation-related biases that prevent accurate or otherwise
meaningful nutritional interpretation. Moreover, food refuse deriving from ritual uses may
or may not have been eaten.
Beginning in the late 1970s, a collaboration between an archaeologist and a geochemist on
the analysis of stable isotope composition of carbon in archaeological human remains (van
der Merwe & Vogel, 1978; Vogel & van der Merwe, 1977) started a revolution in dietary
reconstruction having implications for issues of interest in the study of past populations, all
linked either directly or indirectly to diet. Following this pioneering research, stable isotopes
have become a standard data set for addressing fundamental questions about diet and dietary
adaptation in past populations, enhancing our ability to characterize past human diets and
their variation in general (Burton, 2008; Katzenberg, 2008; Lee-Thorp, 2008; Schoeninger,
2010). Simply, the documentation of isotope signatures passed from the foods being eaten to
the consumer allows the identification of diet with considerably greater precision than with
conventional archaeological data involving recovery of plant and animal remains alone.
These biochemical signatures do not represent a “reconstruction” of diet; rather, they facilitate
the identification of consumption profiles of different foods eaten by past populations. If
consumption profiles are identified accurately, then it becomes possible to get at the real
“meat” of the matter, namely, nutritional inferences and broader implications for understand-
ing past human adaptation in a broad biocultural context.
Regional and methodological perspectives on isotope and elemental analysis in archaeo-
logical remains and identification of dietary signatures are presented in a number of compre-
hensive reviews of the older literature (Katzenberg & Harrison, 1997; Pate, 1994; Price et al.,
1985; Sandford, 1992, 1993a; Schoeninger, 1989, 1995a; Schoeninger & Moore, 1992;
Schwarcz & Schoeninger, 1991) and newer developments building on earlier advances

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
302 Isotopic and elemental signatures of diet, nutrition, and life history

(Ambrose & Krigbaum, 2003; Brown & Brown, 2011; Burton, 2008; Katzenberg, 2008; Lee-
Thorp, 2008; Schoeninger, 2010; Schwarcz & Schoeninger, 2011). This chapter discusses
isotopic and elemental analyses of archaeological human remains and the ways in which
these analyses contribute to our understanding of dietary behavior, nutritional ecology, and
milestone events in early life (e.g., weaning) and in later life (e.g., residential change). The
record is based largely on bone collagen, the structural protein that makes up most of the
organic component of bone, but other sources such as bone mineral, tooth enamel, and
calculus serve as isotope-based dietary proxies (Austin et al., 2013; Schoeninger, 2010; Scott &
Poulson, 2012). Importantly, because enamel does not remodel, the chemical signature
remains largely unchanged once the enamel is fully formed. In contrast, because bone does
remodel over the life course, its signature reflects diet and other aspects of biochemistry for
well over a decade (Hedges & Reynard, 2007; Hedges et al., 2007). Thus, an individual aged
45 years at death has a hard-tissue record of diet from both the juvenile years (teeth) and at
least the last 10 years of his/her life (bone).

8.2 Isotopic analysis


8.2.1 Background
Isotopes are chemical elements that have the same number of protons and electrons, but differ
in the number of neutrons. Unlike radioactive or unstable isotopes (e.g., 14C), stable isotopes of
the same element do not transmute over time. Most elements exist in two or more isotopic
forms. Of the several hundreds of stable isotopes across all elements, ten or more elements
have at least two isotopes with biological significance. Of these, two elements in particular
have received the most attention by anthropologists in interpreting earlier diets, namely
carbon (C) and nitrogen (N). In addition, stable isotopes of oxygen (O), hydrogen (H), and
radiogenic strontium (Sr) provide important insights into diet and contribute important data
from which to reconstruct individual and population residential mobility.

8.2.2 Stable carbon isotopes


12
C and 13C
Carbon has two stable isotopes, 12C and 13C. Field and laboratory studies involving controlled
feeding experiments have shown that stable carbon isotope ratios in an animal’s tissues,
including bone, reflect the ratios in diet (Ambrose & Norr, 1993; Bender et al., 1981; DeNiro &
Epstein, 1978; Howland et al., 2003; Jim et al., 2004; Schoeninger, 2010; Tieszen et al., 1983).
The relative abundance of these isotopes between dietary resources is quite small. Thus, the
ratios in tissues are expressed in parts per thousand (read as parts “per mil” or ‰) relative to
an international standard (marine fossil, Belemnitella, from the Peedee geological formation
in South Carolina [PDB]), as delta (δ) values. These values denote differences that originate in
plant photosynthetic pathways, including either C3 (Calvin-Benson), C4 (Hatch-Slack), or
CAM (crassulacean acid metabolism). C4 plants discriminate less against the isotopically
heavier 13C isotope when using CO2, the carbon source for all terrestrial plants, from the
atmosphere. As a result, C4 plants have less negative δ13C values than C3 plants. In temperate
areas, most plants are the C3 variety (e.g., some grasses, trees, shrubs, tubers) and have

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 303

18

16
Marine

14
d15NAIR ( ‰ )

12 Preagricultural
Agricultural
C4-No marine
10

8
C3-No marine

-25 -23 -21 -19 -17 -15 -13 -11 -9 -7 -5


d13CPDB ( ‰ )

Figure 8.1 Conceptualized representation of preagricultural and agricultural diets on


coastal Georgia. C4-based (maize) diets of agriculturalists are significantly enriched in
13
C relative to preagricultural diets, which consist primarily of marine and C3 plant
components. (From Schoeninger et al., 1990; reproduced with permission of authors and
the American Museum of Natural History.)

δ13C values averaging −26‰, with a range of −22‰ to −38‰ (Tieszen, 1991). In C4 plants,
plants that are typically adapted to hot and dry climates (e.g., tropical grasses like maize,
sugar cane, some amaranths, chenopods, setarias), the reduced discrimination against the
13
C isotope results in values averaging −12.5‰, with a range from about −9‰ to −21‰
(Tieszen, 1991). On average, C4 plants have δ13C values that are about 14‰ less negative than
C3 plants and their consumers (Figure 8.1). Plants with CAM photosynthesis pathways (many
cacti and succulents) have δ13C values that overlap the values in C3 and C4 plants, because
they use either a C3 or C4 pathway as determined by environmental circumstances.
The human consumers of these plants retain the differences in δ13C values, but with values
shifting approximately 1‰–4‰ in experimental settings (DeNiro & Epstein, 1978; Hare et al.,
1991; Schoeninger, 2010; Tieszen & Boutton, 1989; Tieszen et al., 1983) and 5‰ in field
investigations (Lee-Thorp et al., 1989; Schoeninger & Moore, 1992; van der Merwe, 1982)
from the food to what is observed in their bone collagen. Sometimes called the “fractionation”
factor, this offset value represents what is better termed the “diet-to-tissue spacing” or
“trophic effect,” because fractionation technically refers to specific equations that measure
the relative amount of two stable isotopes before and after a particular chemical reaction
(Katzenberg, personal communication; and see Fry, 2006). Photosynthesis is a well-known
chemical reaction, but there are a number of reactions that occur between consumption of a
particular food and the formation of bone collagen (or apatite). Moreover, fractionation in
specific reactions is a constant and it is known that diet-to-tissue spacing is not a constant.
Collagen requires essential amino acids for its formation, and therefore reflects the majority
(~60%) of the δ13C protein component of diet. On the other hand, the mineral carbonate from

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
304 Isotopic and elemental signatures of diet, nutrition, and life history

bone apatite represents more of an approximation of the whole diet, which may include
carbohydrates and fats, in addition to protein not used for protein tissue synthesis (Krueger &
Sullivan, 1984; and see Ambrose & Norr, 1993; Ambrose et al., 1997; Cooke et al., 1996;
Harrison & Katzenberg, 2003; Tieszen & Fagre, 1993).
Marine plants have δ13C values that are between the values of C3 and C4 terrestrial plants
due to variation in their carbon sources, ranging from detritus of terrestrial origin (mix of
local terrestrial plants), dissolved CO2 (−7.0‰), and dissolved carbonic acid (0‰) (Schoenin-
ger & Moore, 1992). Thus, marine organisms consuming these plants have values ranging
from closer to C3 plants at one extreme to C4 plants at the other. Marine fishes and mammals
have δ13C values that are less negative (by about 6‰) than do animals feeding on C3-based
foods, and values that are more negative (by about 7‰) than do animals feeding on C4-based
foods (Schoeninger & DeNiro, 1984).

Maize agriculture in regional perspective


The importance of maize in diets of past populations in North America was the first case to
demonstrate the essential nature of stable isotope analysis in bioarchaeology. In non-coastal
settings or regions of the world where the confounding influence of marine foods on carbon
isotopic signatures is not present (see later), the introduction and increased use of domesti-
cated C4 plants have been documented with remarkable precision by stable carbon isotope
analysis. Beginning with Vogel & van der Merwe’s (1977; van der Merwe & Vogel, 1978) study
of prehistoric and early historic populations from eastern North America, regional and site-
specific studies reveal the first use of and increased reliance on maize – the only major
economically important C4 plant used by native societies in this area (Figure 8.2). In prehis-
toric Ontario, isotopic values derived from collagen begin at a low of −20.5‰ before AD 1000,
indicating a C3-based diet (Harrison & Katzenberg, 2003; Katzenberg, 1993; Katzenberg et al.,

-5

-10
Mean d13 C‰ (±1 s.d.)

-15

-20

-25
-4000 -3000 -2000 -1000 0 1000
Age (BC/AD)

Figure 8.2 Temporal change in stable carbon isotope ratios in eastern North America. Error
bars indicate one standard deviation from the mean where N>1. (Adapted from Ambrose,
1987; courtesy of the Center for Archaeological Investigations. © 1987 by the Board of
Trustees, Southern Illinois University.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 305

-5

d13 C‰ -10

-15

-20

-25
2500 BC 1000 BC 500 BC AD 1 AD 500 AD 1000 AD 1500
Date

Figure 8.3 δ13C values from bone collagen for each sample (circles) documents an increase
in C4 plant consumption after AD 1000 in Ontario, Canada. Mean values from the same date
are indicated by filled circles and mean values are indicated by squares. (From Harrison &
Katzenberg, 2003; reproduced with permission of authors and Elsevier Ltd.)

1995; Schwarcz et al., 1985; van der Merwe et al., 2003). After AD 1000, however, isotope
values rapidly become considerably less negative (Figure 8.3). The presence of relatively less
negative δ13C values in later prehistoric and historic sites in these settings is generally
consistent with findings from other localities in the interior of eastern North America
(compare with Ambrose, 1987; Boutton et al., 1991; Buikstra, 1992; Buikstra & Milner,
1991; Schurr, 1992; Schurr & Schoeninger, 1995; and see following discussion). Analysis of
apatite values shows somewhat earlier use of maize in the diet beginning before AD 1000
(Harrison & Katzenberg, 2003; Katzenberg, 2006). These values suggest that maize was likely
eaten, but not in appreciable amounts. Overall, this record shows that while maize began to
enter diets of native populations in Ontario by AD 500, it did not become a staple until after
AD 1000, when there was a clear and quick shift from C3- to C4-dominated subsistence.
For the broad region of midwestern and interior northeastern North America, the period AD
500–1300 evinces a trend of increase in δ13C values from the earliest to latest periods, but with
notable variation, related to geography, climate, and perhaps cultural preference. In the
southern region of the eastern United States (states of Missouri, Arkansas, Tennessee), the
Ohio River valley, and the American Bottom of the Mississippi River valley, isotope analysis
indicates appreciable increases in maize consumption that occurred rapidly but late in
prehistory (Boutton et al., 1991; Buikstra et al., 1988; Greenlee, 2006; Reber, 2006). In
contrast, populations living in some areas (e.g., Illinois) have gradually higher δ13C values
that suggest a slower rate of adoption of maize (Buikstra et al., 1987).
Temporal comparisons of Ontario populations show a trend that is intermediate between
the southeastern US and Illinois samples (Katzenberg et al., 1995). Katzenberg and coworkers
(1995) suggest that the variation between northern and southern regions of eastern North
America reflects a shorter growing season in the former than the latter, and hence, a relatively
reduced use of maize. They speculate that the later full-blown adoption of maize in the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
306 Isotopic and elemental signatures of diet, nutrition, and life history

southern area may reflect the existence of an established agricultural lifeway at the time of the
introduction of maize (e.g., squash and other indigenous plants). As is characteristic of other
cultural phenomena, dietary practices may have been adopted quickly by some groups and
not by others for reasons that remain unclear. Additionally, other factors influencing stable
isotope values include regional variation in δ13C values of maize, consumption of C3 versus C4
plants by economically important animals (e.g., deer), fish consumption practices, and the
presence of other C4 plants in the diet (e.g., amaranth in the middle Mississippi and Ohio River
valleys) (Katzenberg et al., 1995).
Even within relatively restricted geographic regions, there are apparent spatial differences
in maize consumption. For example, maize appears to have been a greater part of diet in the
outlying areas around the late prehistoric Mississippian center of Cahokia than in Cahokia
itself (Buikstra & Milner, 1991). These differences may represent social distinctions between
the core (elite) and hinterlands (sub-elite) of the Cahokia chiefdom. Moreover, the somewhat
more negative δ13C values in high-status individuals than in low-status individuals within the
Cahokia center point to slightly lower consumption of maize in those elite that have greater
access to a more diverse diet and animal sources of protein (Ambrose et al., 2003).
Analysis of δ13C values in the late prehistoric Ohio River valley drainage similarly
suggests a link between social complexity and maize production and consumption (Schurr &
Schoeninger, 1995). Comparison of populations from late prehistoric Fort Ancient sites
(Baum, Gartner, Feurt, SunWatch) with contemporary populations from organizationally
more complex Mississippian sites (Angel, Wickliffe Mounds) reveals very different isotopic
compositions. Average δ13C values are generally less than −11% for the Fort Ancient samples,
whereas the values are generally more than −10% for the Mississippian samples. Additionally,
there is a strong correlation between site size and isotope values in both this series of sites and
other Fort Ancient and Mississippian period sites. These and other observations (e.g., location
of Mississippian sites in areas with higher agricultural productivity [compare with Ward,
1965]) suggest that, on average, maize played a more important role in subsistence strategies
of Mississippian chiefdoms than of Fort Ancient groups. Schurr and Schoeninger (1995)
speculate that Fort Ancient groups had a subsistence system that was more stable than that
practiced by Mississippian groups, as indicated by the persistence of Fort Ancient but not
Mississippian societies into the historic period. This persistence may reflect the fact that Fort
Ancient populations developed in situ and were well adapted to the limitations of the Ohio
Valley ecosystem. Mississippian societies may represent a cultural intrusion into the region
and practiced an agricultural and subsistence system that may not have been as well suited to
this setting. These complex societies did not share the relatively greater long-term success of
the Fort Ancient cultures – they effectively migrated out of this area and/or became more
similar in structure to the Fort Ancient type of adaptation well before the arrival of Europeans.
Overall, this setting shows a confluence of complex interactions and movements between
regions.
In the late prehistoric central Mississippi River valley in the American Bottom and north of
the American Bottom, spatial analysis of δ13C values shows considerable variation, even
within small regions. Some individuals had C3 diets, whereas others had C4 diets to varying
degrees (Hedman et al., 2002; Rose, 2008; Schoeninger, 2009). The variation does not relate to
sex, age, or status, and appears to reflect resource extraction that is highly geographically
localized, suggesting that food consumption practices, while having broad temporal patterns

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 307

(i.e., maize as a staple in the last several centuries prior to European contact), are immediate –
simply, travel distance to acquire food is limited. Similar patterns of localized dietary
regimens are also well documented in the Georgia Bight region (Hutchinson, 2002; Hutch-
inson et al., 1998; Larsen, Griffin et al., 2001; Norr, 2002; and see later). These findings
suggest the possibility that multiple subsistence strategies were practiced, perhaps reflecting
group or individual decisions pertaining to risk management (Schoeninger, 2009; Winter-
halder & Kennett, 2009).
On the other hand, the local variation described here may represent cultural differences in
foodways. For example, in the rich agricultural region represented by the Ohio River valley
region, analysis of stable carbon isotope ratios reveals key differences between the Fort
Ancient and Mississippian period cultures. The traditional perspective has been that they
are two distinctive cultural traditions (Griffin, 1943). However, in order to explain the sharing
of attributes, some argued that this was due to migration of groups of Mississippians and
interregional contact (Griffin, 1967; but see Pollack & Henderson, 2000). Cook and Schurr
(2009) have revisited these differences in cultural, social, and dietary attributes of Fort
Ancient and Mississippian traditions, especially in spatial and temporal contexts. Rather than
showing a dichotomy in diet in the comparison of Fort Ancient and Mississippian sites,
comparisons of carbon isotope ratios reveal increasing similarity of diets over time. During
the later years, the Fort Ancient and Mississippian stable isotope ratios are statistically
indistinguishable. This confluence of diet in these late prehistoric societies occurs simultan-
eously with the presence of Mississippian architecture and ceramics in the Fort Ancient region
(Cook & Schurr, 2009). Cook and Schurr (2009) make a strong case that the growing similarity
of diet represents the migration of Mississippians into the Fort Ancient region. Thus, at least
with regard to this regional pattern, when viewed in temporal perspective along with a rich
archaeological context, dietary variation may be due to social dynamics involving migration
and interaction.
In addition to documenting variable patterns of maize consumption in the broad region of
the Eastern Woodlands, these studies point to another important finding. Pollen and other
non-isotopic evidence indicate that maize was likely eaten by native populations quite early
in some areas of the Eastern Woodlands (e.g., 3500 yBP at Lake Shelby, Alabama [Fearn & Liu,
1995]). Isotopic analysis of a wide array of prehistoric native populations indicates, however,
that maize did not become economically important and serve as a key source of energy until
very late in prehistory, certainly after AD 900 in eastern North America (Figure 8.2). As
discussed earlier, in the Great Lakes region, analysis of apatite suggests that maize was
consumed well prior to late prehistory. However, populations were not committed to maize
production as a dominant component of diet. This temporal pattern documenting the increas-
ing focus on maize coincides with other evidence of increased morbidity, increased warfare,
and declining skeletal robusticity in later prehistory (Cohen & Armelagos, 1984; Larsen,
1990b; Larsen & Milner, 1994; Steckel & Rose, 2002).
Stable isotope analysis presents a less clear picture of maize consumption in western North
America, primarily because native populations consumed significant amounts of various
foods with δ13C values similar to maize, such as bison meat, amaranth, and cacti (Schoeninger &
Moore, 1992). For example, less negative δ13C values (−8‰) were identified in skeletal series
from northern Texas and Oklahoma (Habicht-Mauche et al., 1994). These values reflect the
consumption of cactus and other nondomesticated C4 plants, bison, and perhaps some grown

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
308 Isotopic and elemental signatures of diet, nutrition, and life history

or traded maize. In other late prehistoric Great Plains settings (e.g., Crow Creek, South
Dakota), less negative values also represent the consumption of a mixture of C3 and C4 diets,
although maize probably contributed significantly to native diets (Bumsted, 1984).
Analysis of carbon isotope ratios in the American Southwest and eastern Great Basin points
to temporally distinctive patterns of dietary variation. Comparison of early (AD 1275–1330)
with late (AD 1330–1400) populations at Grasshopper Pueblo, Arizona, reveals a trend toward
less negative δ13C values, indicating an intensification of maize agriculture (Ezzo, 1993). This
increase coincides with a period of resource stress, decrease in dietary diversity, and environ-
mental disruption associated with climate change during the later occupation of Grasshopper
Pueblo. Ezzo (1993) contends that these disruptions and a decreasing quality of diet may have
contributed to the abandonment of the region in the late fourteenth to early fifteenth
centuries. This pattern of intensification of maize consumption is also present in the Pueblo
period (AD 750–1350) in the Anasazi region of southwestern Colorado (Coltrain et al., 2007).
This gradual increase began when local populations first became reliant on maize at about
400 BC (Coltrain et al., 2007).
At the Pecos Pueblo site, New Mexico, where subsistence practices involved intensive maize
farming (Kellner et al., 2010; Kidder, 1932; Spielmann, 1991; Spielmann et al., 1990),
the comparison of the pre-Plains trade period (AD 1200–1450), Plains trade period
(AD 1450–1600), and historic (mission) period (AD 1600–1838) revealed that there were no
temporal changes in stable carbon isotope ratios (Kellner et al., 2010). Average δ13C values of
−7.7‰, −7.6‰, and −7.7‰, respectively, document a consistent diet for the 400-year period
based on heavy maize dependence and consumption across the population (no differences in
sex or age).
While maize was a dietary staple at Grasshopper and Pecos Pueblo, in the Great Salt Lake
region of the eastern Great Basin, isotopic analysis reveals that Fremont populations experi-
enced a dramatic reduction in C4 consumption and reversion to exploitation of wild resources
in later prehistory. That is, for the period of c. AD 400–1150, there is an increase in
δ13C values, expressing a shift and increased commitment to maize farming (Coltrain &
Leavitt, 2002; Coltrain & Stafford, 1999) (Figure 8.4). As with the Southwestern population,
these eastern Great Basin groups relied on maize as an important component of diet and
subsistence economy. However, unlike the Southwestern populations, there is a sharp drop in
δ13C values (to −18.1‰), revealing the complete shift to a foraging adaptation in later
prehistory. After AD 1150, native populations abruptly and completely eliminated a food
source that heretofore had been a central part of their diet. While various factors were
involved in this rapid shift, analysis of climate data derived from tree rings and pollen profiles
shows a marked reduction in summer rainfall and truncation of the growing season (Coltrain
& Stafford, 1999; Grayson, 2011). This profound climate impact rendered the continuation of
large-scale maize production impossible, and native populations greatly curtailed agricultural
productivity. This climate shift, combined with a reduction in rainfall, contributed to an
adaptive response involving a return to foraging.

Farming in Latin America


Isotopic analyses of Mesoamerican and South American populations also show a highly
variable pattern of C4 plant (maize) consumption. In the Tehuacan Valley, Mexico, maize
consumption appears far earlier than in other regions of North America, as early as 4000 BC

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 309

-10

-12

-14
d13 C‰

-16

-18

-20

400 600 800 1000 1200 1400


Calibrated Date AD
GSL Wetlands Warren/Willard

13
Figure 8.4 δ C values from bone collagen samples exhibit diverse plant consumption from
AD 400 to AD 1150. After AD 1150, Great Salt Lake region populations ate more C3 plants,
shown in more negative δ13C values. (From Coltrain & Leavitt, 2002; reproduced with
permission of authors and the Society for American Archaeology.)

(DeNiro & Epstein, 1981; Farnsworth et al., 1985). In contrast, analysis of stable carbon
isotope ratios in lower Central America reveals that maize use was very minor or altogether
missing in Panama for the same time period (5000–3000 BC) and was adopted considerably
later in Costa Rica (AD 300–1550) (Norr, 1991).
Analysis of prehistoric and contact-era Mayan skeletons from Belize, Honduras, and
Guatemala provides an important picture of dietary variation in Mesoamerica. The Maya
lowlands contain a wide diversity of plants and animals with potential dietary importance,
but poor preservation of food remains – plants and animals – greatly limits understanding of
foodways in this otherwise well-documented region. It has long been recognized that maize
was an important food source in past Mayan populations. Mayan skeletal series show an
abundance of skeletal pathology indicating poor-quality nutrition and elevated disease
burdens. Populations consuming the greatest proportion of maize have an especially elevated
prevalence of skeletal stress indicators, including dental caries, enamel hypoplasia, and cribra
orbitalia/porotic hyperostosis (see Chapters 2 and 3) (Scherer et al., 2007; White, 1997;
Whittington & Reed, 1997; Wright, 2006; Wright & White, 1996). Due to the paucity of
subsistence-related data from archaeological localities, it is difficult to link these indicators of
health to any one particular subsistence strategy.
The picture of dietary ecology in the ancient Maya is informed by a robust record of stable
isotope variation. Temporal and spatial comparisons of δ13C values from environmentally
heterogeneous locations of Lamanai and Pacbitun, Belize, suggest that prehistoric Mayans
placed less emphasis on maize than did some other Mesoamerican groups (e.g., Tehuacan)
(White & Schwarcz, 1989; White et al., 1993, 1994). Nonlinear temporal changes suggest

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
310 Isotopic and elemental signatures of diet, nutrition, and life history

variable emphasis on maize in the region. Pre-Classic populations (1250 BC–AD 250) at
Lamanai have relatively high δ13C values (−12.4‰) at both sites, suggesting strong reliance
on maize. Late and Terminal Classic Lamanai, however, have increasingly negative
δ13C values, indicating a decrease in C4 consumption (White, 1997; White et al., 1993). In
contrast to Lamanai, the record from Pacbitun shows an increased reliance on maize begin-
ning in the Preclassic and peaking in the Late Classic periods (White, 1997; White et al., 1993).
These elevated levels of C4 diets in the Late Classic are similarly displayed at Copán (Reed,
1994, 1999; Whittington & Reed, 1997) and at inland sites in Petén, Guatemala, such as
Piedras Negras and Altar de Sacrificios (Scherer et al., 2007; Wright, 1997, 2006).
Following the collapse of the Classic period Maya in the eighth to ninth centuries AD,
δ13C values at Lamanai (Pacbitun was abandoned by this time) show a marked reversal in the
trend of decreased maize consumption documented for the Late and Terminal Classic periods.
Post-Classic and Historic period Mayan skeletons have substantially less negative values
(−9.3‰ and −9.9‰, respectively), indicating an increased reliance on maize. White and
Schwarcz (1989) interpret this trend as representing a doubling of maize consumption in less
than a century. Thus, in late prehistoric and early contact-era Maya, maize consumption was
high, a finding similar to estimates of 65% to 86% of diet based on maize in Maya Indians in
contemporary society (White et al., 1994).
In addition to the temporal trends, populations with access to marine resources show lower
δ13C values than interior populations having limited or no access to marine foods (Gerry &
Krueger, 1997; Scherer et al., 2007). For example, Lamanai, a coastal population, has a
considerable amount of variation in δ13C ratios, with some individuals consuming relatively
less maize than others. In contrast, in the interior, Copán populations show less variation and
generally high levels of commitment to maize across the community, and consequently higher
δ13C ratios. This pattern is strikingly similar to analyses showing regional variation in coastal
and interior settings and with the localized nature of diet focusing on extraction of resources
that are nearby (and see earlier discussion).
In the entire record of dietary patterns for the Maya region, little evidence for an association
between social collapse and dietary change can be provided (White, 1997; Wright, 2006;
Wright & White, 1996). Based on isotopic analysis, prehistoric diets appear to have been
linked in a general sense with major events in Maya history, but the variation is so great that
no argument can be made tying resource stress – as indicated by dietary change – to the
collapse and abandonment of the southern Maya lowlands (Wright & White, 1996).
A growing record of past diets based on stable isotope variation is available from South
America, especially in the Peruvian Andes. In order to track key aspects of dietary
variation in diverse ecological and cultural settings in this region, Tykot and coworkers
(2006) documented the record at Pacopampa in highland northern Peru and Cardal and
Tablada de Lurin in the Lurin River valley of the central coast of Peru. These sites give an
indication of dietary variation for the Early Horizon period of the first millennium BC, a
time during which maize has traditionally been regarded as a staple crop. Predictably,
analysis of stable carbon and nitrogen isotopes reflect a highly diverse set of diets in this
complex setting. Overall, the record shows that, like coastal populations of Peru and
Ecuador (Ericson et al., 1989; Tomczak, 2003; Tykot & Staller, 2002; van der Merwe
et al., 1993), there is a strong commitment to marine foods with a later increase in maize
in diets.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 311

-3.0
Material
Bone apatite
Tooth enamel

-5.0

-7.0
d13 C‰

-9.0

-11.0

-13.0
~
Guanape Salinar Pre-structural Structural Post-structural
Gallinazo Gallinazo Gallinazo
Figure 8.5 Bone apatite and tooth enamel δ13C values increase from the Guañape to
Post-Structural Gallinazo periods in Cerro Oreja, Peru. These dietary changes correlate
with transitions in sociopolitical complexity up to and after state formation in the
Gallinazo phase. (From Lambert et al., 2012; reproduced with permission of authors
and the Society for American Archaeology.)

In later coastal settings, maize production and consumption intensified over time and
played a crucial role in the rise of complex societies serving as an important precursor to
state development. In the Moche Valley of north coastal Peru, for example, Lambert and
coworkers (2012) tested the hypothesis that intensified maize production preceded the
appearance of the complex political economy fueling the state system associated with the
Early Intermediate period (c. 400 BC–AD 600). Diachronic analysis of bone apatite and tooth
enamel from the Cerro Oreja site documents dietary changes associated with the rise and
establishment of increasing sociopolitical complexity and possible state-level organization
(Figure 8.5). During this time frame, there appears to have been no change in δ13C from the
Guañape phase (c. 1800–400 BC) to the Salinar phase (c. 400–1 BC). This period of dietary
stasis was followed by progressive and swift enrichment of 13C in the Gallinazo phase
(AD 1–200) by about 2.5‰. The change in isotope ratios coincides with a marked increase
in dental caries in the final period representing the state system (Gagnon & Wiesen, 2013).
The record overall supports the hypothesis that intensified production of this storable crop
was likely an important antecedent to the rise of the Moche state during the AD 200s or
300s. Similar peaks representing maize intensification are also documented in the rise of
later state systems, such as the Inca Empire (Burger et al., 2003; Hastorf, 1990; Hastorf &
Johannssen, 1993).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
312 Isotopic and elemental signatures of diet, nutrition, and life history

Maize consumption and social status


Comparisons of δ13C values from high- and low-status prehistoric Central and South Ameri-
can Native Americans suggest differences in maize consumption along social lines, and they
are in patterns consistent with expectations derived from analysis of dental caries and stature.
At Pacbitun and Altun Ha, social differences in maize consumption are indicated by the
presence of higher carbon isotope values (more maize) in high-status individuals than (less
maize) in low-status individuals (White, 1997; White et al., 2001, 2006) (Figure 8.6). The
reverse is indicated at Lamanai, where high-status individuals consumed less maize than did
low-status individuals (Coyston et al., 1999; White & Schwarcz, 1989; White et al., 1993), a
pattern similar to Caracol (Chase & Chase, 2001) and Piedras Negras (Scherer et al., 2007). This
suggests that at Lamanai, high-status adults may have had greater access to protein derived
from C3 consumers and had better diets generally than low-status individuals. High- and low-
ranking individuals at Copán express no differences in δ13C mean values (Reed, 1994;
Whittington & Reed, 1997). However, high-status individuals at this Mayan center show a
greater range of variation in δ13C values than do low-status individuals, indicating perhaps
greater dietary variability and access to more food choices in the elite (Reed, 1994; Whitting-
ton & Reed, 1997). White (2005) hypothesizes that the choice of an elite diet is related to its
ease of access – wherever maize is more difficult to produce (e.g., central lowlands), it will be
more highly valued than wherever maize is easy to produce. Food in this setting is also
heterarchically used by elites to create social identity both within communities (White et al.,
2001) and between neighboring communities (Metcalfe et al., 2009). In addition, status may
have been not only more regionally differentiated, but also less noticeable during the Post-
Classic and Historic periods (Williams et al., 2009).

15
d15 N‰

PIT
10

URN CRYPT/CIST

5
-15 -10 -5
d13 C‰

Figure 8.6 Carbon isotope ratios from bone collagen vary among individuals from urn,
pit, and crypt/cist burials at Pacbitun, Central America. These distributions indicate that
high-status individuals (crypt/cist burials) consumed more C4 plants in their diet than
lower-class individuals (urn and pit burials). (From White et al., 2006; adapted and
reproduced with permission of authors and Elsevier Ltd.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 313

In the La Florida site in the northern highlands of Ecuador, carbon isotope analysis shows
distinctive differences between high- and low-status individuals during the Chaupicruz
phase (AD 100–450) (Ubelaker, 2000; Ubelaker et al., 1995). Elite individuals recovered from
elaborate tombs having emerald beads, marine shells, textiles, and other exotic items have less
negative carbon isotope values than low-status individuals (δ13C ¼ −10.3‰ and −11.6‰,
respectively). Ethnohistoric and archaeological (mortuary) evidence suggests that these dif-
ferences may reflect a greater consumption of maize beer in high-status adults, especially
males. Moreover, if high-status individuals were consuming more tortillas and other maize-
based food than were low-status individuals, the caries prevalence would be similar in the two
social groups (Ubelaker, 2000).
The dietary signatures of carbon isotope values and inferences drawn from dental caries
prevalence provide congruent results in comparisons of high- and low-status individuals. At
Copán and Lamanai, dental caries prevalences are substantially higher in low-status adults
than in high-status adults, and very high-status adults have relatively low δ13C values. Both
data sets indicate greater consumption of maize in low-status individuals (White, 1994). At
Copán, the greater range of values and lower caries prevalence in high-status individuals
suggest a more varied, less carbohydrate-rich diet (Reed, 1994).

Sex differences in diet and inferences of gendered foodways


For the Maya, adult males show less negative δ13C values than adult females at Pacbitun
(White, 1997; White et al., 1993, 2006), Altun Ha (White et al., 2006), Dos Pilas (Wright, 1997),
Copán (Reed, 1994; Whittington & Reed, 1997), and the Terminal Classic period for Altar de
Sacrificios and Seibal (Wright, 1997). In the Pasión region (represented by Altar de Sacrificios,
Seibal, Dos Pilas, Itzan, Aguateca, Tamarindito, and Petexbatun sites), there is a general
pattern of higher δ13C values in males than females, especially in the Terminal Classic period
(Wright, 2006). These findings point to a restructuring of diet involving gender distinctions
and changes in how food was distributed across this complex social landscape of the Maya
lowlands. That is, males consumed more maize than did females in these settings.
Although a completely different setting, but with an emphasis on maize and increasing
social complexity, the eastern Great Basin Fremont population also shows greater consump-
tion of maize in males than in females. While daily dietary intake and differential access to
protein in the comparison of the sexes may have been the key factor for explaining this
unusual pattern, Coltrain & Leavitt (2002) make the case that the more important factor that
serves to distinguish female and male diets is the ritual function of diet, and especially
consumption of maize in the form of beer. If this is the case, then it suggests that patterns
of isotopic variation in this setting were driven by ritual behavior (White, 2005). In contrast,
analysis of maize farmers from Utah Basketmaker II setting of the Four Corners region of the
American Southwest reveals somewhat greater consumption of maize in females than males
(Figure 8.7). However, what is especially striking about the series is the strong distinction in
δ15N values, with females being relatively more depleted than males. Given that cottontail
rabbits are significantly more depleted in δ15N than artiodactyls (deer, bighorn sheep), the case
can be made that women in this broad setting were consuming more rabbits as a protein
source than were males, who likely consumed relatively more artiodactyl sources of protein
(Coltrain & Janetski, 2013; and see following discussion of nitrogen stable isotopes).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
314 Isotopic and elemental signatures of diet, nutrition, and life history

10.0

Female Male
9.0

8.0
d15N‰

7.0

6.0

5.0

4.0
-12.0 -11.0 -10.0 -9.0 -8.0 -7.0 -6.0
d13C‰

Figure 8.7 Stable nitrogen and carbon isotope ratios show dietary distinctions between
males and females in the pre-Puebloan American Southwest. Slight variability in carbon
isotopes suggests differential maize consumption between sexes. Even more profound is the
dichotomization in meat consumption between sexes: males have more enriched nitrogen
ratios than females, which suggests the consumption of deer/sheep by the former and of
small game (e.g., rabbits) by the latter. (From Coltrain & Janetski, 2013; reproduced with
permission of authors and Elsevier Ltd.)

Lamanai females and males show no difference in stable carbon isotope values, suggesting
a similarity in their diets (Coyston et al., 1999; White, 1997; White & Schwarcz, 1989). Reed
(1994, 1999) and White and coworkers (1993) contend that sex differences in diet at Pacbitun
and Copán represent variation in socioeconomic status, with males consuming more maize
than do females. With regard to Copán, Reed observes that this difference “parallels the
observation of higher frequencies of anemia, infection, and a statistically significant higher
rate of caries in females than in males” (1994:216). These findings of reduced consumption of
maize in females relative to males contradict the results presented from analysis of these other
data sets. For example, higher caries prevalence in females than males suggests greater dietary
emphasis on maize-based foods among women than among men in this setting. Thus, while
both isotopic and caries analyses point to possible differences in diet by gender, the contra-
dictory results do not provide a clear picture of what these differences may have been. At
Lamanai, like the isotope evidence, dental caries prevalence data suggest distinctive differ-
ences in diet between females and males (White, 1994).
The variable nature of diet in comparing isotope and dental caries suggests that these data
sets may reflect somewhat different aspects of diet. Dental caries is caused by acids produced
by bacterial metabolism of carbohydrates (sugar). Carbon isotope data derived largely from
collagen – the basis for most of the Mayan isotopic investigations – provides information on

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 315

protein consumption as well as diet generally (Ambrose & Norr, 1993; but see White et al.,
2001 advocating use of carbonate). Thus, changing patterns of diet – especially with regard to
the consumption of maize – will not necessarily be expressed in the same way in the analysis
of dental caries and stable carbon isotope ratios. The general congruence of temporal trends in
caries prevalence and in carbon isotopes in the Maya lowlands with other regions argues for a
generally overlapping picture of these data sources and dietary shifts. For example, caries
prevalence declines dramatically in the Terminal Classic, but then peaks in the Post-Classic,
indicating respective decrease and increase in maize consumption. This pattern is identical to
dietary changes indicated by δ13C values (White, 1988; White et al., 1994). The interpretations
of dietary patterns are enhanced by use of both direct (isotope) and indirect (caries)
approaches.

Maize consumption by Euroamericans


Soon after the arrival of the first European explorers to the Americas in the late fifteenth
century, an exchange of a wide range of foods, including plants, began between the Old World
and the New World. From the Americas, maize was introduced to and widely adopted by
Europeans, not for consumption by humans but as food for domestic farm animals (Hawke &
Davis, 1992). As Europeans settled the Eastern Woodlands of North America, and especially as
they pushed westward into the American Midwest (the so-called Corn Belt) in the nineteenth
century, maize was quickly recognized for its productive potential. Historical sources suggest
that maize continued to be viewed by many early Euroamerican settlers as a food for farm
animals, unfit for human consumption. Thus, maize was economically important, but not for
direct human use; Old World C3 plants introduced to the Americas – wheat, barley, and rye –
remained of greater direct dietary importance.
Analysis of Euroamerican skeletons from pre-twentieth-century cemeteries indicates that
maize was variably used throughout the Eastern and Midwestern regions of the United States
and Canada. Isotopic analysis of the remains of the Cross family, an early nineteenth-century
pioneer family in Illinois, shows a narrow range of δ13C values (−14.1 to −11.3‰) around a
mean of −12.4‰ (Larsen, Craig et al., 1995). The high degree of homogeneity in values is to be
expected in a temporally restricted, closely related group of individuals living in the same
locality and eating the same foods on a daily basis. The high value indicates the likely
importance of maize in this setting, and is similar to prehistoric Native Americans living in
the same general region (e.g., Dickson Mounds, −10.8‰; Schild, −12.3‰; Norris Farms,
−12.6‰ [Buikstra & Milner, 1991]) and other areas of the Eastern Woodlands (Buikstra,
1992; Katzenberg et al., 1995). The Cross family δ13C values are also consistent with other
evidence of dietary practices and specific crops grown by the family. Probate records enu-
merate a crop of approximately 20 acres of maize on the Cross homestead. Although this
maize likely served as food for farm animals, isotope evidence indicates that it was also an
important source of food for the family. The consumption of domestic farm animals eating
maize may have also contributed to these high values in the human samples.
These isotope values from Illinois show some similarities, but also some contrasts, with
other Euroamerican samples analyzed from other contexts in the Eastern Woodlands.
Determination of δ13C values from remains of the nineteenth-century Harvie family in
southern Ontario reveals very little variation in diet (Katzenberg, 1991a). Unlike the Illinois
series, the Harvie stable carbon isotope values are quite different from late prehistoric

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
316 Isotopic and elemental signatures of diet, nutrition, and life history

maize-dependent native populations living in the region prior to European contact (Katzen-
berg et al., 1995; Schwarcz et al., 1985). The mean δ13C value is −18.7‰, which is similar to
fifteenth- to seventeenth-century European values of −18‰ to −19‰ (compare with
B. Kennedy, 1989). These findings suggest, therefore, that the Harvie family did not eat maize,
but rather, mostly C3 plants, such as wheat, barley, rye, and oats. The Harvie dietary
composition is much more in line with conventional interpretations of nineteenth-century
foodways: farm animals ate maize; farmers did not.
Analysis of nineteenth-century US military buried at the Snake Hill site, Ontario, indicates
a wide range of isotopic variation in contrast with the Cross and Harvie pioneer families
(Katzenberg, 1991b). The mean value from the Snake Hill series is −15.8‰, with a range of
−18.5‰ to −12.5‰. This variation reflects the presence of individuals recruited from all over
the northeastern United States and the high degree of variability in regional cuisines,
especially with regard to use of C3 (e.g., wheat and rice) and C4 (maize) plants. Thus, it is
not surprising that these values range between a mostly C3 diet and a mixed C3/C4 diet
(Katzenberg, 1991b).

Old World C4 plant use: millet


In the archaeological past, virtually all European plant domesticates with economic signifi-
cance (e.g., wheat, barley, and rye) were C3 plants. An important exception is broomcorn
millet (Panicum mileaceum), the only major C4 plant. Millet, a tropical grass, is present in
central Europe perhaps by the late fifth millennium BC (Murray & Schoeninger, 1988), but is
not a strong signal of diet isotopically in humans until the Iron Age, beginning c. AD 800
(Bonsall et al., 2004; Le Huray & Schutkowski, 2005; Murray & Schoeninger, 1988) and
adopted or continuing in historic times (Lightfoot et al., 2012; Reitsema & Vercellotti, 2012;
Reitsema et al., 2010). As with maize in later contexts, millet has been assumed to have been
primarily used as food for farm animals. Analysis of collagen from Iron Age skeletons from
Magdalenska Gora, Slovenia, indicates less negative δ13C values, in sharp contrast with
virtually all other non-coastal European samples from various time periods (compare with
B. Kennedy, 1989). Because marine foods were likely not part of the diets of this group,
consumption of C4 plants or animals consuming C4 plants best explains these values
(Murray & Schoeninger, 1988). It is unlikely that C4-consuming animals contributed appre-
ciably to Iron Age diets in Slovenia because isotope values derived from animal remains are
very negative, unlike the values determined from human remains. Similarly, Bronze Age
skeletons from northern Italy revealed enriched 13C, indicating a strong commitment to millet
production and consumption (Tafuri et al., 2009). In contrast, contemporary skeletons from
southern Italy yielded a record of low δ13C, indicating a C3 diet. Consumption of millet in
these settings may also have been linked to social status, with millet representing a low-status
food. At Medieval Trino Vercellese, for example, high-status males consumed less millet than
did low-status males, and the former show less pathology and stress indicators than the latter
(Reitsema & Vercellotti, 2012) (Figure 8.8). Interestingly, no status distinctions occur in
females, a finding consistent with dental caries and other health indicators. Moreover, the
diets of low-status females were similar to those of high-status females and high-status males.
This record for females suggests that women of reproductive age may have benefited from a
cultural buffer involving better diets (Reitsema & Vercellotti, 2012).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 317

13.0
Male, high status
12.0 Male, low status
Female, high status
11.0
Female, low status
10.0
d15N‰

9.0

8.0

7.0

6.0
-20.5 -19.5 -18.5 -17.5
d13C‰

Figure 8.8 Dietary differences are distinguishable between sexes and statuses at Medieval
Trino Vercellese, Italy. Although high-status males demonstrate less consumption of C3
millet plants than do low-status males, no significant discrimination in diets is apparent
between high- and low-status females. (From Reitsema & Vercellotti, 2012; reproduced
with permission of John Wiley & Sons, Inc.)

Millet became a central part of the diet in northern China by 9000 BC (Barton et al., 2009;
Pechenkina et al., 2005; van der Merwe, 1992; Yang et al., 2012). Analysis of δ13C values
indicates that millet contributed well over 50% of carbon in the diets of some groups. The
record of first domestication is situated in the Yellow River (Huang He) basin, initially with
incipient and short-term use and later involving intensive production and consumption by
humans and as fodder for domesticated pigs and dogs, and perhaps chickens. Analysis of
stable carbon isotope ratios in a number of sites reveals remarkably high δ13C values
(Pechenkina et al., 2005), but also with considerable variation, even within a tightly restricted
geographic area. The implications of these heavily millet-based diets are similar to those
settings where maize became a central component of diet (North and South America). That is,
millet is an incomplete protein, is among the poorest sources of protein, and is especially
deficient in the amino acid lysine. Anemia and other physiological disorders occur in these
kinds of dietary regimens. In this setting, relatively little evidence of porotic hyperostosis is
associated with early farming populations (Pechenkina, Benfer et al., 2007). Indeed, at the
Jiangzhai site (6900–6000 yBP) where δ13C values are relatively high, no individuals dis-
played lesions that are associated with anemia. However, enamel hypoplasias and other
lesions representing growth disruption and compromised health are highly prevalent (Pechen-
kina, Benfer et al., 2007).
From the earliest exploitation in China, millet cultivation spread into the Eurasian steppes.
In the Minusinsk Basin in southern Siberia, comparisons of earlier and later cultures reveals
a rapid uptake in millet at c. 1400 BC or slightly later (Svyatko et al., 2013), a shift that
took place within a century or less. The speed of adoption is remarkable, and is analogous
to the fast adoption of maize in some New World settings (Larsen, Griffin et al., 2001;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
318 Isotopic and elemental signatures of diet, nutrition, and life history

Larsen, Schoeninger et al., 1992; and the next section) or the shift from marine to
terrestrial foods in Neolithic Europe (Tauber, 1986; and following). On the other hand, when
millet is adopted in the Minusinsk Basin, it occurs in a highly variable way, involving
some individuals consuming it and others not. The adoption of millet in this setting occurred
simultaneously with the adoption of the horse and increasing social stratification
in southern Siberia (Svyatko et al., 2013). Therefore, this disparity may reflect variation in
social status in this setting.
Along with millet, sorghum – also a C4 domesticate – dominated the archaeological record
for much of Nubian prehistory (White & Schwarcz, 1994). Analysis of bone and other tissues
shows that although these plants were present in substantial amounts, diets were based
principally on C3 cultigens (e.g., wheat and barley) (White & Schwarcz, 1994), but analysis
along the length of hair shafts from mummies indicates seasonal shifts from C3 to C4 staples
(White, 1993). Isotopic evidence indicates a shift toward consumption of more C4 plants in the
X-group (AD 350–550), a period characterized by political instability, alterations in trade
patterns, and decreased water availability as the Nile River lowered. In the following Christian
period (AD 500–1400), more negative isotope values reveal a decline in C4 plant consumption.
This shift in diet appears to have accompanied an increase in elevation of the Nile and
generally improved economic conditions.

Marine diets and coastal settings: the foraging-to-farming transition


In coastal areas where no C4 plants are consumed, stable carbon isotope data provide an
important means of assessing the relative importance of marine and terrestrial foods. In
Scandinavia, clear shifts in dietary orientation are documented in the comparison of Meso-
lithic and Neolithic populations. In coastal Mesolithic Danish populations, for example,
generally less negative δ13C values (mostly between −17‰ and −13‰) indicate a reliance
on marine foods, and increasingly so over time (Tauber, 1981, 1986; and see Bennike &
Alexandersen, 2007; Fischer et al., 2007; Jørkov et al., 2010; Richards, Price et al., 2003)
(Figure 8.9). In later coastal groups (Neolithic, post-Neolithic), δ13C values are more negative
(−19‰ to −21‰), indicating a striking shift to and complete dependence on terrestrial C3
foods, such as from domestic plants and from farm animals, meat or milk, or both. Isotopic
evidence indicates that post-Mesolithic Danish and Swedish populations consumed few
marine foods, despite close proximity to them, and were remarkably similar, regardless of
place or time (Jørkov et al., 2010; Lidén, 1995; Lightfoot et al., 2011; Richards, Price et al.,
2003; Tauber, 1981, 1986).
Elsewhere in Europe, a similar pattern emerges in comparing isotope values from earlier
foragers (Mesolithic) and later farmers (Neolithic and post-Neolithic). In particular, coinciding
with the shift to farming, there are rapidly decreasing δ13C values in Britain and Ireland
(Richards, Schulting et al., 2003; Schulting, 2011) and France (Herrscher & Le Bras-Goude,
2010; Schulting & Richards, 2001), illustrating a strong reliance on marine foods in the
Mesolithic that is replaced by a strongly terrestrial diet in the Neolithic, a pattern that
continues in northern Europe as well as in other European settings through prehistory and
into historical times (Craig et al., 2009; Fuller et al., 2003, 2010; Kjellström et al., 2009; Le
Huray & Schutkowski, 2005; Lubell et al., 1994; Müldner & Richards, 2007a, 2007b; Polet &
Katzenberg, 2003; Reitsema et al., 2010; Richards et al., 2002, Richards, Price et al., 2003;
Schutkowski et al., 1999; Yoder, 2010; but see Kosiba et al., 2007; Lightfoot et al., 2011;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 319

16

14

12
d15N‰

10

RIA MED
6 VIKA MES
NEO
4
-22 -21 -20 -19 -18 -17 -16 -15 -14 -13 -12 -11
d13C‰

Figure 8.9 Diachronic dietary changes in Denmark: RIA (Roman Iron Age), VIKA (Viking
Age), NEO (Neolithic), MED (Medieval), and MES (Mesolithic). Decreases in δ13C over time,
from the Mesolithic to the Neolithic and subsequent periods, demonstrate an increase in
13
C-depleted marine foods. (From Jørkov et al., 2010; reproduced with permission of
authors and John Wiley & Sons, Inc.)

Prowse et al., 2004). In northern England, a significant return to use of marine foods is
observed only after the rise of deep sea fishing, c. AD 1000 (Müldner & Richards, 2007a,
2007b). These analyses reveal general dietary patterns, geographic and temporal, but they also
show how specific diets can be within restricted areas (Herrscher & Le Bras-Goude, 2010;
Polet & Katzenberg, 2003; and see Hutchinson et al., 1998 for a New World setting).
The overall record shows that the foraging-to-farming (Mesolithic to Neolithic) transition in
Europe was not a gradual process, but was rapidly achieved in many of the settings where
there is a substantial isotopic record of diet. Even coastal populations saw some adaptive
advantage to abandoning – either wholly or mostly – the use of marine foods in diet and
deriving protein and energy from terrestrial foods. This also implies a general reduction in the
quality of diet and a more homogeneous diet relative to Mesolithic forager-fisher-gatherers
(Schulting, 2011). Moreover, in the Baltic region where agriculture is adopted later in the
Neolithic, there is evidence for a greatly reduced marine component as terrestrial domesticated
resources become emphasized (Lillie & Budd, 2011).
The quick shift in diet is not restricted to northern Europe but it is also well documented in the
Mediterranean region, indicating that the earlier record of the agricultural transition taking
place in this setting was also rapid. In particular, terminal Neolithic populations from Alepo-
trypa Cave, coastal southern Greece, have very negative δ13C values, for both collagen (mean ¼
−19.9‰) and apatite (mean ¼ −13.1‰) (Papathanasiou, 2001, 2003, 2011; Papathanasiou
et al., 2000). These findings indicate that the diets of these groups were largely focused on
terrestrial C3 plants and animals with near abandonment of marine resources. The narrow range
of values (collagen: −19.5‰ to −20.2‰) indicates a remarkably high degree of homogeneity in
diets. Moreover, this dietary adaptation is associated with a marked increase in stress indicators
(e.g., porotic hyperostosis) (Papathanasiou, 2001, 2003, 2011; Papathanasiou et al., 2000),
similar to that found in other regions of the globe (e.g., Mesoamerica).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
320 Isotopic and elemental signatures of diet, nutrition, and life history

Correspondingly, an abrupt shift from marine to terrestrial food resources in the compari-
son of Mesolithic and Neolithic populations (c. 8500–4500 yBP) in Portugal is indicated by
isotopic analysis (Lubell et al., 1994). Values for the samples drawn from various coastal and
near-coastal sites range from −15.3‰ to −20.4‰. The more negative values are predomin-
antly Neolithic and the less negative values are Mesolithic. The presence of less variability in
the Neolithic period suggests increasing homogeneity in diet during the period where plant
and animal domesticates were more heavily relied upon (Lubell et al., 1994). Unlike the
pattern of increased prevalence of dental caries in New World settings with the increasing
reliance on plant domesticates, there is a marked decline in dental caries prevalence, number
of carious tooth surfaces, and premortem tooth loss in permanent mandibular molars in late
Mesolithic and Neolithic Portuguese in comparison with earlier populations from the region.
The relatively high prevalence of dental caries in the Mesolithic period is probably related to
the consumption of cariogenic plants such as nondomesticated figs (Lubell et al., 1994; and
see Chapter 3).

8.2.3 Stable nitrogen isotopes


14 15
N and N
Field and laboratory feeding studies demonstrate that stable nitrogen isotope ratios in an
animal’s tissues, including bone, reflect ratios of diet (DeNiro & Epstein, 1981; Hare et al.,
1991; Wada, 1980). The ratios determined from bone samples are expressed as ‰ relative to
the international standard of atmospheric nitrogen (N2, or Ambient Inhalable Reservoir, AIR).
The organism’s stable nitrogen ratio in its tissues is largely determined by its trophic level in
the local food web. Simply, as nitrogen derived from foods is incorporated into the organism’s
tissues – including skeletal and dental – bonds between 14N and another element (e.g., carbon
or oxygen) breaks down at a more rapid rate than those containing 15N. As a result, the lighter
14
N isotope is excreted with urea (Schoeninger & DeNiro, 1984). The result is the tissues of the
consumer organism have a higher δ15N value (more positive ratio) than what has been
consumed (Drucker & Bocherens, 2004). Two broad groups of plants derive their nitrogen
by alternative means, depending on the source of nitrogen. Legumes derive nitrogen from
bacteria (Rhizobium) located in the roots of the plant. Metabolic processes of the bacteria
allow it to use atmospheric nitrogen, which is 0‰ in combining the nitrogen with hydrogen,
oxygen, and other elements, thus providing the nitrogen to the plant via the roots. Other
plants, however, derive their nitrogen from organic material that has decomposed or is in the
process of doing so in the soil surrounding the plant roots. This decomposition process
produces soil with nitrogen (e.g., ammonia [NH3] or nitrate [NO3]) that is higher in 15N than
is atmospheric nitrogen because relatively more 14N is released into the atmosphere (e.g.,
nitrous oxide) during the process of decomposition. This natural fertilizing process is also
enhanced via “manuring” or organic chemical fertilizers. The former is likely an ancient
practice (Bogaard et al., 2013; Finucane, 2007), whereas the latter is used from the nineteenth
century. Non-legumes that take up natural fertilizers or manure are relatively more
15
N-enriched, and therefore have somewhat higher δ15N values (~3‰). Most synthetic
fertilizers have values closer to 0‰ because the process uses atmospheric nitrogen. In
terrestrial contexts, herbivorous animals have δ15N values that are about 6‰ and
carnivorous animals have δ15N values that are about 9‰. These values are general, and can

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
Figure 2.5 Rickets in left juvenile (4–5 years) ulna (left) and radius (right);
Medieval period, Gruczno, Poland. (From Kozłowski, 2012; photograph
by Tomasz Kozłowski.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(a)

Figure 2.7a Porotic hyperostosis on parietals; Santa Maria de los Yamassee,


Amelia Island, Florida. (Photograph by Mark C. Griffin.)

(b)

Figure 2.7b Histological section from Figure 2.7a. The linear orientation and general
morphology of diploic cavities are consistent with iron deficiency anemia. (Schultz et al.,
2001; reproduced with permission of authors and University Press of Florida.)
Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 2.8 Cribra orbitalia; Santa Catalina de Guale de Santa Maria, Amelia Island, Florida. (From Larsen,
1994; photograph by Mark C. Griffin; reproduced with permission of John Wiley & Sons, Inc.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(a)

(b)

Figure 2.9 Abnormal bone growth and porosity in glabellar region (a) and on greater wing of
sphenoid (b), diagnostic of scurvy; Chornancap, Peru. (Klaus, 2013; reproduced with permission of
author and Elsevier Ltd.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at

(a) (b)
Enamel

Perikymata

Striae of
Retzius
(c)
Dentin
Occlusal border
of defect

Occlusal wall
perikymata
(stress episode)

Cervical/occlusal wall junction


cervical wall perikymata
(recovery period)
Pulp
Cervical border
of defect

Figure 2.12 Cross‐section of tooth displaying the relationship between striae of Retzius (a) and perikymata (b). For full caption see page 45.
Figure 2.13 Maxillary dentition showing enamel hypoplasia on incompletely erupted central incisors;
anatomical specimen. (From Larsen, 1994; photograph by Barry Stark; reproduced with permission of
John Wiley & Sons, Inc.)

Figure 2.20b Actual cross section of compact bone showing major microstructures discussed in text.
For full caption see page 61.

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 3.1 Mandibular carious lesions; Ochsenfurt, Germany. Root caries have affected the
cement-enamel junction of the molar (right) and resulted in the complete destruction of the
tooth crown (left). (Photograph and copyright by Leslie L. Williams; reproduced with
permission.)

Figure 3.6 Periosteal reaction on mid-diaphysis of adult tibia; Ochsenfurt, Germany.


(Photograph by Leslie L. Williams.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(a)

Figure 3.3a Antemortem tooth loss of mandibular dentition; Ochsenfurt, Germany.


(Photograph and copyright by Leslie L. Williams; reproduced with permission.)

(b)

Figure 3.3b Edentulous individual; Santa Catalina de Guale Santa Maria, Amelia Island,
Florida. (From Larsen, 1994; photograph by Mark C. Griffin.)
Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
1

100 mm

Figure 3.8 Confocal laser scanning image of archaeological bone. For full caption see page 93.

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 3.11 Development deformity of anterior mandibular dentition (enamel
hypoplasia and Hutchinson’s incisors) associated with congenital syphilis.
For full caption see page 100.

Figure 3.13 Destructive lesions on thoracic vertebral bodies pathognomonic of tuberculosis;


Lambayeque, Peru. (Klaus 2010; reproduced with permission of author and Elsevier Ltd.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 3.14 Destructive lesions on bodies of thoracic and lumbar vertebrae, resulting in kyphosis of the
spine (Pott’s disease, tuberculosis); Norris Farms, Illinois. (Photograph by George R. Milner and the
Illinois State Museum.)

Figure 3.15 Tuberculosis rib lesions, Terry Collection. (Photograph by Charlotte A. Roberts.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(a)

Figure 3.16a Rhinomaxillary syndrome in leprosy; Early Medieval, Lauchheim, Germany.


(From Boldsen, 2008; reproduced with permission of author and John Wiley & Sons, Inc.)

Figure 4.1 Perimortem cranial depressed fracture at the glabella of the frontal bone; Phum
Snay, Cambodia. (Domett et al., 2011; reproduced with permission of authors and Antiquity.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 4.9 Cut marks on adult frontal (scalping); Norris Farms, Illinois. (From Milner &
Smith, 1990; reproduced with permission of authors and Illinois State Museum.)

Figure 4.6 Antemortem diaphyseal fractures


of right radius (right) and left ulna (left);
Çatalhöyük, Turkey. (Photograph by Scott
Figure 4.4 “Parry” fracture of right ulna. In this
Haddow.)
case, the fracture site resulted in a disunion, or
pseudoarthrosis; Lambayeque, Peru.
(Photograph by Haagen Klaus.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 4.7 Atrophy and healed fracture of right humerus; Shanidar,
Iraq. The left humerus is normal. (Photograph and copyright by Erik
Trinkaus; reproduced with permission.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(a)

Figure 4.10a Arrow point in adult sternum; Norris Farms, Illinois. (From Milner & Smith,
1990; reproduced with permission of authors and Illinois State Museum.)

(b)

Figure 4.10b Projectile point in left tibia; Norris Farms, Illinois. (From Ryan & Milner, 2006;
reproduced with permission of authors and Elsevier Ltd.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 4.16 Pacatnamú mass burial, second layer of human skeletal remains; Jequetepeque Valley, Peru. (Verano, 2008; reproduced with
permission of author and Springer Verlag.)
Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(a)

(b)

(c)

(d)

Figure 4.18 Composite distribution of cut marks associated with ritual killing events. For full caption
see page 153.

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 4.19 Sharp force trauma to face associated with a bladed weapon; Towton, United
Kingdom. (From Novak, 2000; reproduced with permission of author and Oxbow Books.)

Figure 4.21 Incomplete trephination and depressed skull fracture on right frontal bone;
Cinco Cerros, Peru. (Verano, 2007; reproduced with permission of author and University
of Arizona Press.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(a)

(b)

Figure 4.20 Cut adult humerus (a) and scapula (b); Tatham Mound, Florida. (From Hutchinson & Norr,
1994; reproduced with permission of authors and John Wiley & Sons, Inc.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(a)

(b)

Figure 5.1 Pathognomonic indications of osteoarthritis at various joints: deformation of shoulder


joint visible on both right scapula (left) and humerus (right) (a); new bone growth, marginal osteophytes,
and eburnation on humerus (left), ulnae (center), and radius (right) (b); marginal osteophytes and pitting
at wrist (c); eburnation and osteophytosis on femoral knee surface (d); pitting and osteophytosis of
cervical (e) and lumbar (f) vertebrae; Morropé, Peru. (From Klaus et al., 2009; reproduced with permission
of the authors and John Wiley & Sons, Inc.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(c)

(d)

Figure 5.1 (cont.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(e)

(f)

Figure 5.1 (cont.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 5.2. Distal right humerus showing eburnation (osteoarthritis); anatomical specimen.
(From Larsen, 1987; photograph by Barry Stark; reproduced with permission of Academic
Press, Inc.)

Figure 5.3. Collapsed thoracic vertebrae; Cochiti, New Mexico. (Photograph by Clark
Spencer Larsen.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 5.4. Schmorl’s depression on superior surface of thoracic body;
Lambayaque, Peru. (Photograph by Sam Scholes.)

Figure 7.2 Summary of craniofacial changes in Nubia, comparing Mesolithic foragers with
Meroitic-Christian farmers (dashed line).

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(a)

(b)

Figure 5.8 Entheseal changes on the medial epicondyle of the humerus (a), radial tuberosity
of the radius (b), and attachments for infraspinatous and teres major muscles.
(Photographs by Sébastien Villotte.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 7.4 Superior (top) view of large mandibular torus; Kodiak Island, Larsen Bay, Alaska,
pre-AD 1000. (Photograph by G. Richard Scott.)

Figure 7.6 Lingual tilting of mandibular first molar; Golovin Bay, Alaska, protohistoric. (Photograph
by G. Richard Scott.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Foraging to farming

Reduced toughness of food

Altered chewing

Reduced masticatory
muscle activity

Altered craniofacial growth

Reduced Reduced size and


facial growth configuration of
masticatory muscles

Reduced Decreased stimulation


face and jaws of bone growth in
face and jaws

Reduced robusticity of
face and jaws

Response of cranial vault and


reduced amount of bone

Shorter, wider, more globular


vault with less-projecting
face and less room for teeth
(increased malocclusion)

Figure 7.3 Masticatory-functional model of craniofacial change in Nubia. (From Larsen,


2014. © 2014, 2011, 2008 by W. W. Norton & Company, Inc. Used with permission of W. W.
Norton & Company, Inc.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(a)

(b)

Figure 7.8 Lateral views of tooth wear in Nubian A-group agriculturalist (a) and Eskimo hunter-gatherer
(b), showing greater angle of wear plane in the former than the latter. (From Smith, 1984; reproduced
with permission of author and John Wiley & Sons, Inc.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
(a)

(b)

Figure 7.9 Occlusal view of tooth wear in Nubian X-group agriculturalist (a) and Mesolithic
hunter-gatherer (b) showing cupped occlusal wear on first molars in the former but not the latter.
(From Smith, 1984; reproduced with permission of author and John Wiley & Sons, Inc.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 7.10 Adult mandibular dentition showing occlusal surface grooves; Great Basin,
Nevada. (Reproduced with permission of John Wiley & Sons, Inc.)

Figure 7.11 Lingual wear on maxillary incisors; Çatalhöyük, Turkey. (Photograph


by Scott Haddow.)

Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 7.12c Confocal microscope 3D image ( 100) of same location on same tooth. (Image courtesy of Melissa Zolnierz and Christopher Schmidt.)
mm

2.5

1.5

0.5
3

0
Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
Figure 7.13 Chipped mandibular incisors from the Cathedral of Santa Maria in Vitoria,
Spain, c. AD 1400–1500. (Photograph courtesy of G. Richard Scott.)

11 168

Yoruba

56 8

3 8 138 3

18 57% 51 90% 43% 10% 12

Algonquin Han

6 65 16 21 32 8%

Zapotec 1 Karitiana Aleutian Chipewyan

16 92%

8 2 13

East Greenland Inuit West Greenland Inuit Naukan

Figure 9.7 Gene flow, reconstructed through modern day single nucleotide polymorphisms,
from Asia to America suggests two migrations occurring after the initial migration into the
Americas: first migration (blue), second migration (green), and third migration (red). Dotted
lines represent mixture events, and solid circles denote hypothesized ancestral populations.
(From Reich et al., 2012; reproduced with permission of authors and Nature Publishing
Group.)
Downloaded from http:/www.cambridge.org/core. University of Warwick, on 11 Dec 2016 at 02:34:28, subject to the Cambridge Core terms of use, available at
http:/www.cambridge.org/core/terms. http://dx.doi.org/10.1017/CBO9781139020398.015
8.2 Isotopic analysis 321

TERRESTRIAL MARINE

Carnivores
(n = 30)

Primary
Carnivores (n = 26)
(n = 6)

Herbivores
(n = 21)

Precision of
isotopic analysis

Plants
(n = 32)
(n = 136)

0 +50 +100 +150 +200


d15N‰

Figure 8.10 Terrestrial and marine plants and animals are represented in broad trophic levels
according to δ15N values in bone collagen. Each trophic level is associated with substantial
enrichment of δ15N. Nitrogen ratios, in turn, can be compared (site-specifically) to these
trophic models to determine primary dietary protein contributions. (From Schoeninger &
DeNiro, 1984; reproduced with permission of authors and Elsevier Ltd.)

vary considerably, depending on the setting. Omnivores, such as humans, have values that are
mostly between those of herbivores and carnivores (Minagawa & Wada, 1984; Schoeninger &
DeNiro, 1984; Schoeninger & Moore, 1992). Therefore, this increasing enrichment of 15N as
one moves up the food chain provides the opportunity to determine the trophic level and to
infer the dietary source based on analysis of bone (collagen, in particular) (Figure 8.10). Key to
interpreting these diets, however, is understanding the biochemistry of plants and animals
within the food webs of specific regions (Katzenberg, 2008). Within these specific regions, key
patterns of dietary consumption involving access to protein show important tendencies that
document broad patterns. For example, at Neolithic Çatalhöyük, there are increases in
δ15N values over the site’s 1400-year occupation (Hillson et al., 2013; Pearson, 2013; Pearson
et al., 2013), likely reflecting increasing protein derived from animal sources, peaking in the
terminal phase of the site’s occupation (Figure 8.11), an increase which coincides with the
presence of domesticated cattle and a greater likelihood of the consumption of cattle protein
in day-to-day contexts as well as special feasting (Russell et al., 2013). In contrast, consider-
ably lower δ15N values are documented at Neolithic Nevalı Çori (Lösch et al., 2006), empha-
sizing a different set of local factors – dietary and cultural – that play into understanding
regional patterns of food use. These findings show that communities in some settings
employed food to differentiate between men and women, or the sexes had somewhat different
lifestyles and eating patterns relating to subsistence activities.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
322 Isotopic and elemental signatures of diet, nutrition, and life history

14.0

Female
13.5
Male

13.0

12.5
d15N‰

12.0

11.5

11.0

10.5

10.0
Early Middle Late
Period
Figure 8.11 Stable nitrogen ratios from Neolithic Çatalhöyük increase from the Early to Late
periods, indicating increased dietary consumption of animal meat. (From Hillson et al.,
2013; reproduced with permission of authors and Cotsen Institute of Archaeology.)

Trophic distinctions in marine settings


Owing in part to the presence of zooplankton, zoobenthos, and insect prey at the bottom of
the trophic system in these environments, marine vertebrates often express higher δ15N values
than terrestrial vertebrates (DeNiro & Epstein, 1981; Schoeninger & DeNiro, 1984). In this
regard, δ15N values in terrestrial plants and animals are about 10‰ less positive than in
marine plants and animals (Schoeninger & DeNiro, 1984). Ultimately, differences in marine
and terrestrial environments are reflected in the tissues of the human consumers. These
differences in tissues, including bone, can be used to determine the relative importance of
terrestrial versus marine foods from these ecosystems (Schoeninger et al., 1983; Schwarcz &
Schoeninger, 1991). The distinction is especially well known for continental coastal settings,
whereas in a number of regions where fish derived from lakes or streams provide a potential
source of food, the food webs are quite complex and often difficult to characterize, making
interpretations of δ15N values problematic (Katzenberg & Weber, 2009; Katzenberg et al.,
2012; Lillie et al., 2011; Reitsema et al., 2010).
Nitrogen isotopic signatures are influenced by a number of other factors, especially climate.
Generally, cool forest soils have low δ15N values due to higher nitrogen-fixation and mineral-
ization rates, and hot savannah or desert soils have high δ15N values (Ambrose, 1993;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 323

Schwarcz et al., 1999). Other contexts produce generally high δ15N soils, as in areas with a
history of evaporation (e.g., saline soils) and those enriched in organic materials (e.g., guano).
The very high values reported from analysis of human bone samples in some desert settings
may be explained by these factors (Ambrose & DeNiro, 1986a, 1986b; Aufderheide, Tieszen
et al., 1988; Heaton et al., 1986; Schoeninger, 1995b). In some settings, terrestrial animals
living in arid environments have higher δ15N values than marine animals. Because of climatic
and/or other ecological variables, Late Stone Age coastal and interior foragers in South-
western Cape Province, South Africa, have nitrogen isotope ratios inverse to those expected,
but carbon isotope values clearly identify differences in the exploitation of marine versus
terrestrial resources (Sealy, 1986; Sealy & van der Merwe, 1985, 1986; Sealy et al., 1987). In
other settings, stable nitrogen isotope ratios are useful for identifying consumers of terrestrial
versus marine foods (Schoeninger & DeNiro, 1984; Schoeninger et al., 1983; and see following
discussion).

The life cycle and physiology: breastfeeding, weaning, and the adoption of adult diet
Behaviors impacting the earliest periods of life, not the least of which are the duration of
weaning and the length of time it takes to adopt a complete or near complete adult diet, have a
profound influence on a number of facets of a population’s demographic dynamics. A large
and expanding body of evidence from demographic and anthropological sources, especially
from study of traditional societies, shows that the shorter the duration of reliance on
breastfeeding as the infant’s sole or dominant food source, the sooner the mother can become
fertile again and produce another infant. That is, nursing suppresses ovulation and female
fecundity (Ellison & O’Rourke, 2000; Vitzthum, 1994), and increases the inter-birth interval.
Hypothetically, populations that, on average, have a long-duration breastfeeding regimen will
have fewer offspring than populations with a short-duration breastfeeding regimen. Thus, in
the present world, as in the past, population stasis and growth is very much influenced by
breastfeeding practices as they relate to duration, weaning, and timing of adoption of adult,
non-breast milk diets.
There are many factors that affect fecundity and fertility in a human society (Ellison &
O’Rourke, 2000; Gage, 2010; Vitzthum, 1994). However, all else being equal, breastfeeding
and an infant’s dependence on the mother as the sole source of energy and nutrition, the
duration of this part of early life history, and the time it takes to adopt a wholly adult diet all
play vital roles in shaping demographic characteristics of a population, especially its size and
growth rate. Demographic analysis provides especially important insight into understanding
population growth, stasis, or decline, including the remarkable demographic transition that
accompanied the shift from foraging to farming beginning in the early Holocene. In this
regard, might the dramatic population growth in the Neolithic be explained in part or in whole
by a change in weaning timing?
Buikstra and coworkers (1986) addressed this question by hypothesizing that agriculture
and the development of more efficient ceramic cooking vessels in later prehistory in the lower
Illinois River valley facilitated the production of maize-based gruel. The increased use of this
improved technology for preparing these soft weaning foods thus would shorten the length of
time to transitioning to adult foods. This factor, therefore, might explain the growth in
population in the later intensive farmers compared to the earlier less-intensive farmers (see
Chapter 10). In order to test the hypothesis regarding reduced duration of breastfeeding,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
324 Isotopic and elemental signatures of diet, nutrition, and life history

weaning, and the complete shift to adult foods, Fogel and coworkers (1989, 1997), in their
classic study, documented the δ15N values from fingernails (composed of the protein keratin)
of nursing infants and their lactating mothers over the course of the first year of life. They
found a distinctive trophic effect involving an average increase in δ15N values when compar-
ing the infants to their mothers. When the infants were fully weaned, they exhibited values
within 1‰ of their mothers; within three to five months following cessation of consumption
of mother’s milk, the infant values were identical to those of their mothers. In the interim–the
weaning process – as the milk provided to the infant via breastfeeding is replaced by other
foods, the δ15N values decline.
Fogel and coworkers (1989, 1997) also compared δ15N values in juveniles and adults from
hunter-gatherers and maize farmers from eastern North American archaeological contexts,
predicting that the hunter-gatherers would express a longer duration in early life of elevated
δ15N values than the farmers. The former are represented by Archaic period remains from Eva,
Cherry, and Ledbetter sites in the Tennessee River valley, and the latter are from the historic-
era Sully site in the northern Missouri River valley. The age changes in δ15N followed exactly
what would be predicted: (1) newborns had values strongly similar to adults, reflecting the
fact that the breast milk is derived from the mother’s diet and has isotope values largely the
same as all tissues in the mother; (2) within several months, infant values increased, reflecting
trophic differences between the consumer (the infant) and the consumed (the mother); and (3)
values increase during the period of breastfeeding, and decrease thereafter, reflecting the
weaning process. Thus, the pattern of δ15N values clearly tracks the duration of breastfeeding
and the subsequent weaning process leading to fully adult diets. Contrary to their expect-
ations, however, Fogel and coworkers found no difference in the timing of weaning between
the hunter-gatherers and the agriculturalists. Nevertheless, their study accomplished an
important goal; namely, establishing the essential link between the weaning process and
how the process is documented via δ15N values in early life. Moreover, the values showed that
weaning is not an abrupt event. Rather, these findings point to a gradual reduction in
δ15N values over young childhood rather than a sudden decline, suggesting that the introduc-
tion of foods – generally having lower δ15N values – and replacement of breast milk is a
gradual process. While many anthropologists and other social scientists tend to regard age-at-
weaning as a specific life history milestone marking the complete termination of breast
feeding, their study also confirmed the growing realization of nutrition scientists that
weaning is a process, sometimes of lengthy duration (Herring et al., 1998). In this way, work
by Fogel and collaborators (1989, 1997; and see Fuller et al., 2006; Nitsch et al., 2011; Schurr,
1998) allows us to model and predict life history patterns relating to early life (Figure 8.12).
Children in the twenty-first century are typically weaned from breast milk completely at
some point between the ages of one and five years. However, it is common in developed
countries for weaning to occur considerably earlier (Dettwyler & Fishman, 1992). Documen-
tation of age profiles of δ15N values in relation to the breast feeding of infants is demonstrated
via stable nitrogen isotopic analysis, revealing the variation and timing from diverse arch-
aeological contexts globally. All of these records reveal a strikingly similar pattern in the spike
in δ15N values over the life course, with an increase in values tracking the period of bone
development, a peak representing full dependence on mother’s milk, the period of weaning,
and the assumption of a fully mature diet. Most of the settings where δ15N ratios have been
documented by age from neonate to adult show very similar patterns overall, but with

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 325

11
-16
Exclusive d15N
breastfeeding
d13C
10 -17

Weaning

9 -18

d15N‰ d13C‰

8 -19
Introduction of
solid foods

-20
7

6
0 2 4 6 8
Age (years)
Figure 8.12 The process of weaning, as modeled with stable carbon and nitrogen ratios,
depicts the incorporation of solid foods into a child’s diet followed by gradual weaning
from maternal milk resources. (From Nitsch et al., 2011; reproduced with permission of
authors and John Wiley & Sons, Inc.)

considerable variation in the timing of the weaning process. For example, at Neolithic
Çatalhöyük and nearby Aşıklı Höyük, in southern Turkey, documentation of δ15N values
reveals no enriched isotope values after the age of one year, suggesting that the weaning
process started before 1.5 years or somewhat later (Hillson et al., 2013; Pearson et al., 2010,
2013; Richards, Price et al., 2003) (Figure 8.13). In contrast, the earlier Neolithic site of Çayönü
Tepesi in southeastern Turkey shows a pattern indicating that the weaning process com-
menced by as much as a year later than at Aşıklı Höyük (Pearson et al., 2010). This represents
significant variation within the larger region, reflecting an entirely different timing of and
approach to weaning. Likely, these differences reflect a combination of factors, including
environmental differences (more arid to the west and more temperate to the east) and dietary
differences (narrower range of foods to the west and wider range to the east) (Pearson et al.,
2010; and see Lösch et al., 2006).
Analysis of δ15N values in a range of settings in the Roman Empire reveals similar patterns
in breastfeeding and weaning, but with significant variation, generally supporting the notion
that weaning is locally determined, primarily by environmental and cultural factors. At Kellis
in the Dakleh Oasis, Egypt, the stable isotope record indicates that weaning foods were likely
introduced by about six months and breast feeding cessation occurred no later than three
years of age, the time at which there is no apparent trophic effect (Dupras, 2010; Dupras &
Tocheri, 2007; Dupras et al., 2001). The record from δ15N in bone collagen reveals the general
pattern, but the use of δ13C in the analysis using the comparison of deciduous and permanent

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
326 Isotopic and elemental signatures of diet, nutrition, and life history

(a)
-15
-16
-17
-18
-19
d13C‰

-20
-21
-22
-23
-24
-25
in utero birth 5 10 15 20
Age (years)

(b)
18

16

14

12

10
d15N‰

0
in utero birth 2 4 6 8 10 12 14 16
Age (years)

Figure 8.13 Stable carbon (a) and nitrogen (b) isotope ratios from Neolithic Çatalhöyük show
a decrease in isotope enrichment after about 1.5 years, which indicates that weaning
processes began before this age. (From Hillson et al., 2013; reproduced with permission of
authors and Cotsen Institute of Archaeology.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 327

teeth provides an additional measure of trophic level that is similar to δ15N. That is, there is a
trophic effect in δ13C representing about 1‰ (DeNiro & Epstein, 1978; Post, 2002), also
documented in the comparison of mother and infant pairs (Fuller et al., 2006). There are some
differences between δ15N and δ13C in the rate of return to the mother’s baseline values.
Namely, δ13C appears to return to the mother’s baseline value sooner than δ15N. At Isola
Sacra, the largest port located near Rome, the strong shift in and variation of δ15N and
δ13C values occur between one and two years of age, indicating the time of weaning and the
process of full adoption of adult diets, perhaps by 2.5–3 years (Prowse et al., 2008).
Documentation of dental caries and tooth wear indicates the kinds of high-carbohydrate
foods that were used in the weaning process. For example, calculus deposits are first observed
in individuals older than two years of age, and occlusal surface wear, albeit slight, is present
in deciduous teeth in the one- to two-year age range. Carious teeth are not present until the
age of two years, and clearly point to the minimal level to which infants and young children
are first exposed to cariogenic foods.
On the northern periphery of the Roman Empire, patterns of nitrogen and stable carbon
isotope variation reveal a very similar outcome, with a decline in δ15N values after the age of
two years. Relatively low δ15N values in infants aged 1.5–2 years indicate an earlier age of
complementary feeding in a series from Queensford Farm, England (Fuller et al., 2006). In this
setting, nearly all juveniles have low (adult) values by the age of three to four years. Overall,
therefore, stable isotope values from the handful of Roman period sites and estimated ages of
peak reliance on the mother’s milk, the duration of weaning, and the full adoption of adult
diet are similar and consistent with written accounts of weaning practices during the Imperial
period.
More recent populations from historical contexts in Europe and North America show the
same pattern of age change in δ15N values, involving an initial and fast-paced rise when the
isotopic signal of breastmilk is reflected in recently formed bone, indicated by the highest
values at about one year of age during the peak period of complete dependency of the nursing
infant on the mother’s milk, followed by a decline in values. It is the period of decline that
provides the picture on the timing and duration of the weaning process. In contrast to the
Roman-period populations, a later Medieval series from the Wharram Percy site (c. AD
900–1500), England reveals an earlier and faster decline in δ15N and δ13C values such that
by two years of age, the values are strongly similar to adult values, indicating that weaning
had commenced well prior to the age of two years (Mays et al., 2002; Richards et al., 2002)
(Figure 8.14).
Analysis of δ15N and δ13C values in the an eighteenth- to nineteenth-century urban setting
of Spitalfields, London, England provides important insight into weaning practices, especially
involving a time during which women were more heavily involved in formal employment and
had a considerably greater presence in the workforce in general. Under these circumstances,
women who were employed away from the home were almost certainly likely to either wean
their infants earlier than in prior times or not breastfeed at all. As a result, mothers were
increasingly using artificial means of feeding their infants (Fildes, 1995; Lewis, 2007). The
Spitalfields series represents an affluent population, with individuals interred in a church
crypt between AD 1729 and 1859 (Nitsch et al., 2011). Analysis of nitrogen and stable carbon
isotopes reveals clear age changes representing the key phases of early childhood involving
dependence on breastfeeding and weaning (Figure 8.15). Hypothetically, an urban setting like

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
15

14

13

12

11

d15N‰ 10

8
Fetal Adult
7

4
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
Age (years)
Figure 8.14 Infants from Wharram Percy exhibit enriched stable isotope nitrogen ratios
during breastfeeding years. By two years old, weaning has occurred, as demonstrated by the
decreased δ15N values, which fall within the average adult δ15N range. (From Mays et al.,
2002; reproduced with permission of authors and Antiquity.)
(a)

17
Female
Male
16 Unknown

15

14
d15N‰

13

12

11

10

0 5 10 15
Age (years)

Figure 8.15 Ratios of stable nitrogen (a) and carbon (b) isotopes from post-Medieval
Spitalfields reveal definitive dietary shifts in early childhood associated with breastfeeding
(enriched isotopic values) and weaning practices. The dashed line and bar represent the
average ( 1 standard deviation) isotope ratios for adult females from Spitalfields. (From
Nitsch et al., 2011; reproduced with permission of authors and John Wiley & Sons, Inc.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 329

(b)

-15
Female
Male
Unknown
-16

-17

-18
d13C‰

-19

-20

-21

-22

0 5 10 15
Age (years)

Figure 8.15 (cont.)

this would show minimal breastfeeding based on the age profile of the δ15N values, but this
series shows a majority of individuals having a period of exclusive breastfeeding. On the other
hand, there is considerable variation in values ranging from no or little departure from adult
values in the birth to two-year-old age range to moderate departure. These values suggest that
some infants were breastfed on an irregular basis, and some were likely never breastfed. The
stable carbon isotope record shows a similar pattern supporting the notion for a high degree of
variability with at least some individuals being breastfed. Because many children have known
ages at death, there is a more precise picture of when weaning was complete based on the
stable isotope record. This record shows that weaning was complete in the population between
the age of one and two years. Similarly, the later population from the nineteenth century
St. Thomas’s cemetery in Belleville, Ontario, displays a pattern of stable nitrogen isotope
values increasing in the first two months (δ15N ¼12.1‰), peaking, and then declining by the
age of one year (Herring et al., 1998). By two years old, all individuals had adopted a fully
adult diet (post-age two years, δ15N ¼10.4‰). Like the Spitalfields series, there is considerable
variation in the period of birth to 1.5 years, with some showing no evidence of breastfeeding.
The documentation of breastfeeding, weaning, and duration of weaning leading to adop-
tion of a fully adult diet suggests that dependence on breastfeeding and shortening of the
weaning period certainly occurred in historical, urban contexts in Europe and North America.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
330 Isotopic and elemental signatures of diet, nutrition, and life history

However, the examples discussed here likely do not represent the range of circumstances of
weaning as documented in archaeological settings. That is, the limited number of historical
contexts (Medieval and post-Medieval) have documented relatively recent times, and cer-
tainly not rural localities. The record of weaning as based on stable isotope analysis shows
considerable variation and not discernible trends. In this regard, in prehistoric archaeological
settings, there is a remarkable range of variation, from relatively short duration of total
breastfeeding (e.g., Native Americans: Schurr, 1998; Watts et al., 2011) to long duration of
total breastfeeding (e.g., Maya: Williams et al., 2005; Wright & Schwarcz, 1998; Gaya
confederacy, Korea: Choy et al., 2010). Moreover, consistent with Fogel and coworkers’
(1997) findings, there is no clear relationship between weaning and subsistence economy or
demographic pattern, at least as they are measured in bioarchaeological contexts (Schurr,
1998; Schurr & Powell, 2005). These findings suggest that while it has become possible to
document dietary signatures in young juveniles, providing an important perspective on
breastfeeding, weaning, and adoption of adult diets, the linkage between fertility, population
growth, and weaning is complex. Much has yet to be learned.
The terminus of the weaning process, as documented by the δ15N and/or δ13C curves,
corresponds closely with the rise in morbidity indicators of physiological stress, such as
hypoplasias and circular caries, and elevated mortality (Dupras & Tocheri, 2007; Fuller
et al., 2006; Herring et al., 1998; Lewis, 2007; Pearson et al., 2010; Reed, 1994). Fuller and
coworkers, for example, clearly document the association between decreasing δ15N values and
low δ13C values and the spike in childhood deaths. This is because once weaning processes
commence, infants simultaneously lose the passive immunity and complete nutrition pro-
vided by the mother that had been provided during the period of breastfeeding, thus com-
promising the infant’s immune system. This double jeopardy – weaning and reduced
immunity – explains elevated infant and child illness and mortality, both in the archaeo-
logical past and in a large number of settings around the world today, especially in developing
countries (Lewis, 2007). In the modern world, weaning and other associated health risks (e.g.,
poor diets and exposure to contaminated food) contribute to debilitation, diarrheal disease,
and early death (Motarjemi et al., 1993). The bioarchaeological record shows that this peak age
of risk has considerable antiquity.

Life after weaning: the development and maturation of adult diet


While a rich record of isotope variation and dietary and nutritional inference is available via
bioarchaeological analyses, much of this record concentrates on the juvenile record or the
adult record, but rarely both. The life history of the full range of dietary history from birth
through adulthood reflects, however, the important perspective on population as a whole. In
this vein, analysis of δ15N and δ13C in bone at Dakhleh Oasis, Egypt and early historic South
Korea shows no apparent change in adults (Choy et al., 2010; Dupras et al., 2001). Conversely,
at Isola Sacra, Italy, there is an increase in δ15N and δ13C, but with the most pronounced
increase in δ15N for females. The latter record of age change suggests that older adults may
have been consuming more meat and fish in their diets than younger individuals. If this record
and its interpretation are correct, then it may reflect greater status and better access to protein
in older adults (Prowse et al., 2005). Similarly, δ13C values in older women at Copán,
Honduras are more negative than in younger women, indicating a reduction in maize
consumption with increasing age (Reed, 1999).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 331

Comparison of female average and male average isotopic signatures shows distinctive
trends indicating gender-based dietary differences between women and men in some but not
all settings. In a comprehensive study of age-related dietary changes in Isola Sacra, Italy,
combined δ15N and δ13C values reveal a gender distinction: males have higher average
δ15N and δ13C values than females, for the population as a whole, and within age categories
(Prowse et al., 2005). This record reveals that males were likely consuming more protein
(fish, animal products) and fewer carbohydrates (plants) than were females. Interestingly, the
δ13C values appear to converge in older adulthood with females becoming more like males.
This suggests that older females had greater access to protein in later life. These differences
match the record of status and gender in Imperial Rome. That is, men had considerably
higher status and access to resources than women. Likely, the dietary differences are part of
the status distinctions along gender lines. Similarly, written accounts document the advice
given by physicians, namely that it is preferable to limit womens’ food consumption, and for
them to avoid the so-called “cold and wet” foods, such as fish and meats (Prowse et al.,
2005:9). On the other hand, other stable isotope analysis of two other Imperial Rome series
(Casal Bertone, Castellaccio Europarco) reveals that diet varies along class lines (Killgrove &
Tykot, 2013). That is, higher-status individuals living closer to the center of Rome (Casal
Bertone) consumed more marine sources of protein than did lower-status individuals living
further from the city center (Castellaccio Europarco). Moreover, those living further from the
center ate more C4 food sources, likely millet, than did their periurban counterparts. These
findings show that diet was variable in Rome, and was likely influenced by status.
Patterns of variation in δ15N and δ13C values show generally greater access to animal
protein for males in Roman and Medieval-period England (Poundbury: Richards et al., 1998;
Queenford Farm: Fuller et al., 2006; Fishergate: Müldner & Richards, 2007b), Medieval Poland
(Giecz: Reitsema et al., 2010), early historic South Korea (Choy et al., 2010), the ancient Maya
(Whittington & Reed, 1997), and prehistoric preagricultural and agricultural settings in the
Ohio River valley and tributaries (Schurr & Powell, 2005). In the Ohio River valley region, the
differences were statistically significant only in the maize-based late prehistoric settings from
Sun Watch and Wickliffe sites. The patterns in these diverse settings suggest significant
differences in diet along gender lines, which is also apparent in terms of oral health, especially
caries prevalence (see Chapter 3). In addition, historical accounts document the differences in
gender roles in eastern North America where males hunted while women focused their
attention on horticulture (Swanton, 1946).
Schurr and Powell (2005) offer that while the variation in isotope values – especially for
nitrogen – may well reflect dietary differences, these disparities more likely represent physio-
logical distinctions (i.e., related to pregnancy and lactation) between adult females and males.
However, a physiological explanation does not account for a lack of sex differences in a
number of settings in North America and elsewhere, where females and males are similar in
their isotopic signatures or where females have higher δ15N values than males. For example,
in a comprehensive study of isotopic variability in the Georgia Bight and northern Florida, no
significant sex differences were identified in either δ15N or δ13C values, regardless of setting
(coastal, terrestrial) or period (Hutchinson et al., 1998; Larsen, Griffin et al., 2001; Larsen,
Schoeninger et al., 1992; and see Kellner & Schoeninger, 2008; Kjellström et al., 2009;
Pechenkina et al., 2005; Tafuri et al., 2009; Ubelaker, 2000). This variation is consistent with

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
332 Isotopic and elemental signatures of diet, nutrition, and life history

the conclusions drawn by Schwarcz and Schoeninger (1991) that there is no or very limited
physiological basis for the differences seen in the comparison of adult females and males. On
the other hand, among the ancient Maya, there are clear sex differences, especially in elite
individuals, throughout the sequence from Pre-Classic through the Historic period (White,
2005). Even in the Historic period (Lamanai, Belize) when considerable social disruption and
other influences considerably altered native foodways, males showed greater consumption of
animal sources of protein than did females, as expressed in carbon and nitrogen stable isotope
ratios. This pattern of greater carnivory in males than females expresses clear gender differ-
ences and differential access to socially valued foods.

Other social distinctions


Interestingly, clearer social distinctions in diet are present in some Medieval European
settings, suggesting evidence of differences in diet along gender and status lines (Fornaciari,
2008; Kjellström et al., 2009; Polet & Katzenberg, 2003; Reitsema et al., 2010). For example,
in the Medieval Polish population from Giecz, adult women have much lower δ15N values
than men, and women have a more homogeneous diet than men, suggesting less meat
consumption in women than men (Figure 8.16). Reitsema and collaborators (2010) suggest
that rather than representing social inequality, these differences reflect variation in access
owing to occupation. On the other hand, at the Medieval setting from Sigtuna, Sweden,
there are clear isotopic differences between high- and low-status individuals living in an
urban context, and strong evidence of social hierarchy (Kjellström et al., 2009). In this
setting, men consumed generally more protein from animal sources than did women, and
like Giecz, women’s diets are more homogeneous than men’s diet. Similarly, and not
unexpectedly, the super-elite Medici family in sixteenth- to seventeenth-century Florence
show values of δ15N similar to carnivores, reflecting their diets rich in animal sources of
protein (Fornaciari, 2008).

11

10
d15N‰

8 Males
Females

7
-20 -19 -18
d13C‰

Figure 8.16 δ13C and δ15N values from the Medieval population of Giecz (Poland) express
differential dietary patterns between the sexes. Males consume diets with more nitrogen
isotopically enriched animal protein. (Data from Reitsema et al. 2010.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 333

Life stresses and isotope variation


The tracking of diet in the life course from birth through adulthood makes clear that
physiology has an important impact on bone chemistry. With the publication of bioarchaeo-
logical (White & Armelagos, 1997) and animal (Hobson et al., 1993) studies linking
15
N enrichment and stress, an important line of bioarchaeological chemistry has opened up
opportunities to document and interpret health as it is represented biochemically for nitrogen
and other isotope analyses, including carbon, sulpher, oxygen, and hydrogen (Katzenberg,
2008; Reitsema, 2013; Warinner & Tuross, 2010; White, Longstaffe et al., 2004). This work
underscores the critical links between physiology – normal and abnormal – in how nitrogen
and other isotopes are mediated by the consumer’s physiology via the fractionation process.
For living humans, some of the best documented influences relating to 15N enrichment and
15
δ N values are revealed in links between morning sickness and body weight changes over the
course of pregnancy in living humans (Fuller et al., 2006). In this regard, owing to morning
sickness and appetite reduction following conception, the loss in body weight occasions an
increase in δ15N values, but following recovery of appetite in later stages of pregnancy, body
weight increased and δ15N values increased sharply. These later changes reflect an increas-
ingly positive nitrogen balance and an attendant shift of nitrogen to tissue growth instead of
being excreted with urea. Similarly, various disease states (e.g., liver disease, diabetes, and
osteoporosis) affect isotope balances.
Stress and compromised health clearly have an influence on isotopic signatures in all of the
body tissues, which is displayed in the variation documented in the bones and teeth studied by
bioarchaeologists. The study of bone collagen from Nubian X-group skeletons reveals that
osteoporotic females have higher δ15N values than non-osteoporotic females, especially in the
third and fifth decades of life (White & Armelagos, 1997). This pattern is consistent with
histomorphometric differences between X-group individuals with osteoporotic and normal
skeletal tissue. In this setting, it appears that the differences in δ15N values reflect differences
in urea nitrogen excretion or altered renal processing and clearance of phosphorus and
calcium (White & Armelagos, 1997). In the hot desert setting of Nubia, water stress may
contribute to elevation in δ15N values. Lactation contributes to urea loss, which has been
linked theoretically to enrichment of 15N in bone collagen (Ambrose & DeNiro, 1986b). These
findings suggest that physiological disruption, more so than diet, may be responsible for the
variation in nitrogen stable isotopes (and bone loss) in this population (White & Armelagos,
1997). More generally, it strongly suggests that stable nitrogen isotopes reflect, at least in part,
non-dietary – especially physiological – factors and may serve as an indicator of osteoporosis
and likely other health conditions in past human populations.

8.2.4 Bivariate use of carbon and nitrogen stable isotope ratios: sorting out diets
in coastal environments
In many coastal areas of the world where marine and terrestrial foodways may be difficult to
identify owing to the complexities of local and regional food webs, the simultaneous analysis
of δ13C and δ15N is very informative. This has been especially so for coastal settings involving
maize production and consumption in New World contexts. Where marine foods and maize
were simultaneously consumed by human populations, the overlapping carbon isotope ratios

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
334 Isotopic and elemental signatures of diet, nutrition, and life history

16
Early Prehistoric

14

Late Prehistoric
12
d15N‰

10
Mission

6
-20 -15 -10 -5
d13C‰

Figure 8.17 Diachronic changes in isotope ratios from Early (black circle) and Late (white
circle) Prehistoric to Mission (hatched circle) periods on St. Catherines Island, Georgia,
reveal dietary shifts away from marine resources and toward terrestrially based plants
and animals. (From Larsen, Griffin et al., 2001; reproduced with permission of authors
and University Press of Florida.)

of marine and C4 foods preclude the discrimination of diets, at least with regard to assessing
the relative contribution of marine foods versus maize (Figure 8.1). That is, what might appear
to be a marine diet may simply be mimicking a C4 diet, and vice versa. In order to circumvent
this complication, the use of bivariate plots of δ15N and δ13C values has served to clarify diet
(Brown & Brown, 2011; Cooke et al., 1996; Katzenberg, 2008; Tomczak, 2003). In the Georgia
Bight, archaeological evidence indicates that marine foods continued to be heavily used
throughout prehistory and into the contact period (Larsen, 1982; Thompson & Turck, 2009,
2010; Thompson and Worth, 2012). Maize consumption is largely based on circumstantial
evidence in this region, including changes in settlement (increased population size and
aggregation), increasing social complexity, and increasing morbidity (e.g., dental caries,
periosteal reactions; see Chapters 2 and 3). Owing to extraordinarily poor preservation of
plant remains in late prehistoric archaeological sites, dietary reconstruction based on the plant
record is inconclusive as to the relative importance of maize as a routine food source.
Analysis of collagen samples from prehistoric foragers and farmers and early and late
contact mission native populations alleviates the incomplete and vague picture of diet in this
setting. Isotopic analysis reveals a distinctive temporal trend showing increasingly less
negative δ13C values and less positive δ15N values, both in local settings (e.g., St. Catherines
Island: Figure 8.17) and in the larger region (Figure 8.18) (Hutchinson et al., 1998; Larsen,
Griffin et al., 2001; Larsen, Hutchinson et al., 2001; Larsen, Schoeninger et al., 1992;
Schoeninger, 2009; Schoeninger et al., 1990). This trend indicates an increasing focus on
terrestrial plants (maize) and animals, coupled with a decreasing reliance on marine resources.
This shift commences during the twelfth century AD with the adoption of maize as a dietary
staple, reaching its peak and extreme degree of focus in Spanish mission native populations

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 335

14

13

12

11
N‰

10
15
d

6
-18 -16 -14 -12 -10 -8
13
d C‰
Georgia early prehistoric, coast Florida mission, inland
Georgia late prehistoric, coast Georgia early prehistoric, inland
Florida late prehistoric, coast Florida early prehistoric, inland
Georgia mission, coast Georgia late prehistoric, inland
Florida mission, coast Florida late prehistoric, inland

Figure 8.18 Mean stable carbon and nitrogen isotope values from coastal and inland Florida
and Georgia skeletal series demonstrate the significant change in diets from prehistoric to
mission periods. The mission-period samples (box) show increased consumption of C4
plants relative to earlier periods, along with a narrower dietary range.

inhabiting St. Catherines Island, Georgia, in the early to mid-seventeenth century, and in the
later descendant groups on Amelia Island, Florida (Hutchinson et al., 1998, 2000; Larsen,
Griffin et al., 2001; Schoeninger, 2009). Overall, this temporal record reveals a near abandon-
ment of marine foods in the mission era (post-AD 1600) in the Georgia Bight. Moreover, at
several of the missions, more negative δ13C values and/or more positive δ15N values in elite
individuals (those interred closer to the church altar) than in the non-elite (those interred
further from the church altar) suggest differential access to high quality foods (Larsen, Huynh
et al., 1996; Stojanowski, 2013a, 2013b).
The prehistoric dietary changes documented in the Georgia Bight are, in some respects, also
present in the western Florida panhandle. In this regard, the late prehistoric Mississippian
center at Lake Jackson and Waddells Pond sites have δ13C values representing a significant
maize diet (mean ¼ −13.7‰; Larsen, Griffin et al., 2001). These are inland localities, and as
predicted, there is a non-marine signature of nitrogen. In sharp contrast to the late prehistoric
Georgia coastal localities, late prehistoric coastal and non-coastal stable carbon isotope ratios
for populations from the northern and central peninsula of Florida are noticeably negative,
expressing results that indicate little C4 (maize) consumption. Maize may have been present in
these diets and there was considerable variation in diet, depending on habitat (Kelly et al.,
2006), but the isotope record suggests that maize or other C4 foods were not predominant in
diet, certainly not to the same extent as in most of the rest of eastern North America. However,
as with the Georgia coastal mission populations, other historic settings in Florida, mission and
non-mission, express considerably higher δ13C values, indicating the spread and adoption of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
336 Isotopic and elemental signatures of diet, nutrition, and life history

maize after European contact (Hutchinson, 2006; Hutchinson & Norr, 1994; Hutchinson et al.,
1998, 2000; Larsen, Griffin et al., 2001). The sudden shift in diet in contact-era Florida and
Georgia suggests that dietary change was a largely cultural, economic, and political endeavor
occasioned by the arrival of Europeans and exploitation of native populations in food
production (Worth, 2001). In contrast, the shift to maize in the prehistoric era likely involved
an adaptive strategy.
Additional analysis of collagen from the late prehistoric Irene Mound site on the north
Georgia coast indicates that although maize generally increased in importance, there is a
marked decrease in δ13C values for the period immediately prior to European contact,
suggesting that use of maize temporarily decreased following its initial introduction in the
region (Larsen, Schoeninger et al., 1992). The decline in maize consumption may be linked
with social and environmental disruption in the Mississippian period in this region during late
prehistory. Similar declines in use of maize during the period preceding contact by Europeans
has been documented in at least one other major Mississippian center in the American
Southeast, Moundville, Alabama (Schoeninger et al., 2000). These findings from Georgia
and Alabama point to the highly variable pattern of maize use in eastern North America.
Analysis of carbon and nitrogen stable isotope ratios helps to clarify the relative consump-
tion of marine foods and maize in other coastal settings in the Americas, including New
England (Bourque & Krueger, 1994; Little & Schoeninger, 1995; Medaglia et al., 1990), Gulf
coast Florida (Hutchinson, 2004; Hutchinson & Norr, 1994), coastal North Carolina (Hutch-
inson, 2002; Hutchinson et al., 2007), Panama and Costa Rica (Norr, 1991), Belize (White &
Schwarcz, 1989), and Peru (Tomczak, 2003). Comparisons of inner and outer coastal popula-
tions from North Carolina show that, in contrast to Georgia groups where agriculture was
adopted regardless of location (inner vs. outer coast), the outer coastal North Carolina groups
continued to rely on oceanic and estuarine resources throughout later prehistory, whereas inner
coastal groups adopted maize farming after AD 1000 (Hutchinson et al., 2007). Similarly,
Chiribaya culture (AD 1000–1450) sites representing coastal (San Geronimo), head of the
coastal valley (Chiribaya Alta), near coastal, but with material culture showing little focus on
marine resource extraction (Chiribaya Baja), and middle valley, located some distance from the
coast with material culture indicating terrestrial exploitation (Yaral), provide an important
context for addressing the degree to which populations with access to marine resources
committed to maize agriculture (Tomczak, 2003). Bivariate comparison of nitrogen and carbon
reveals that diets in this setting show a progressive decrease in both δ13C and δ15N values.
Importantly, this variation indicates that diets are determined on the basis of local extractive
strategies with consumption being highly localized. Predictably, the δ15N values show that
populations living closest to the coast consume the highest amount of seafood (San Geronimo).
The proportion of seafood that was consumed by these populations, at progressive points
further from the coast, reaches its lowest at Yaral, where diets included marine foods, but were
dominated by consumption of C4 foods and animals eating C4 foods (e.g., llama).
Even where maize is not utilized in New World settings, the isotopic compositions involv-
ing carbon and nitrogen are not always clear. Walker and DeNiro (1986) showed that the
simultaneous use of carbon and nitrogen ratios serves to clarify the potentially conflicting
signatures of either element alone in marine environments of the Santa Barbara Channel
region. Archaeological and biocultural evidence indicates a shift in dietary emphasis from
terrestrial to marine resources in later prehistory in this region as well as a generally heavier

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 337

emphasis on marine foods on the islands than on the nearby mainland coast or interior.
Isotopic analysis reveals that the carbon and nitrogen ratios in native populations progres-
sively decrease from the islands to the mainland coast and mainland interior, indicating a
strong correlation of diet with geographical location (Walker & DeNiro, 1986). These geo-
graphic distinctions in isotopic signatures show that terrestrial diets were emphasized in
mainland populations, and marine diets were emphasized in island populations. Additionally,
isotope values from older sites on the mainland coast and interior are lower for both carbon
and nitrogen than those from younger sites in the same region. This finding is consistent with
other archaeological and biocultural evidence suggesting that mainland-oriented diets
became more marine-oriented later in time.
In the Bahamas where terrestrial, reef, and deep-ocean habitats were utilized by prehistoric
populations, Keegan and DeNiro (1988) analyzed numerous potential foods from these
settings. Some reef fish have higher δ13C values and lower δ15N values than other ocean fish.
The analysis of carbon and nitrogen isotope ratios in archaeological human samples from this
setting indicates that prehistoric populations fished primarily in seagrass and coral-reef
ecosystems. The most negative δ13C values are in the initial period of occupation; later
populations have less negative values. This temporal shift in stable carbon isotope ratios
suggests that Caribbean populations became increasingly marine-oriented in later prehistory.
In coastal settings in both the western and eastern Pacific Rim, there has long been
confusion over the role of marine and terrestrial foods consumed. For example, in Japan
and the American Northwest coast, while it has long been understood that both marine and
terrestrial foods were consumed by Jomon foragers, the importance of each has been unclear
(Chisholm & Koike, 1999; Chisholm et al., 1992; Minagawa & Akazawa, 1992). Isotopic
analysis of human remains from a series of sites revealed that hunter-gatherer-fishers from
Hokkaido of northern Japan, from Jomon through Ainu periods, showed few temporal
distinctions and with a pattern of carbon and nitrogen isotope ratios indicative of maritime
mammals. Jomon-period individuals from coastal sites on Honshu, however, consumed a
lesser amount of maritime mammals, and instead relied on combinations of plant food,
terrestrial mammals, and fish. In contrast, populations from the inland site of Kitamura (Late
Jomon, 4000 yBP) on the main Japanese island display very negative δ13C and less positive
δ15N values, signaling a largely terrestrial diet. Interestingly, some complexity in these dietary
patterns is observed for the inland populations: individuals from the Boji site (Late Jomon,
c. 4000 yBP) have isotopic signatures that are consistent with the consumption of freshwater
fish (Yoneda et al., 2004).
Although representing a very different ecosystem, some island populations of the western
Pacific that acquired agriculture show a similar temporal pattern of shifts from heavily marine
to heavily terrestrial diet. Non-isotopic archaeological data from a number of settings show
that populations likely adopted agricultural subsistence within the last 2000 years of prehis-
tory, but marine foods remained important dietary targets (Field et al., 2009). In order to
identify dietary change in the southwest Pacific, Field and coworkers (2009) documented
stable isotope ratios for carbon and nitrogen in Fiji, from the period of 2700–300 yBP from
both coastal and terrestrial contexts. The general pattern is clear: the earliest samples from
Waya Island show an isotopic pattern that is fully marine, similar to other Pacific settings,
such as Rota, Guam, and Saipan in the Marianas (Ambrose et al., 1997) and Tutuila in
American Samoa (Valentin et al., 2011). In contrast, the later samples from the Sigatoka

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
338 Isotopic and elemental signatures of diet, nutrition, and life history

Valley display values that characterize a terrestrial adaptation involving domesticated plants.
Overall the record for Fiji reveals a shift from marine to terrestrial (Field et al., 2009; Valentin
et al., 2006, 2011), with the change occurring earlier on larger islands having more arable land
than smaller islands. Interestingly, however, while the general trend toward terrestrialization
of diet is clear, exceptions to the pattern suggest local variation in diet based on access to
marine foods (Jones & Quinn, 2009).
Much of this record documenting coastal environments and resource acquisition shows the
complex nature of dietary interpretation based on the analysis of stable isotope ratios.
However, the use of bivariate plots of stable carbon and nitrogen values alleviated some of
the issues with dietary interpretation. There are many coastal environments worldwide where
use of these plots is quite complex, such as in large freshwater lakes (Katzenberg & Weber,
2009; Katzenberg et al., 2012). For example, in the Lake Baikal region (Siberia, Russian
Federation), Katzenberg and coworkers used stable isotopes to reveal variation in use of
terrestrial mammals versus freshwater marine sources (fish and seals), largely owing to
differences in the stable carbon and nitrogen isotope signatures. Adding to the complexity
of the setting are findings relating to isotopic signatures of freshwater fish. The analysis of
Katzenberg and coworkers (2012) reveals considerable variation. However, a number of
patterns emerge, including important distinctions between specific settings that provide a
means to show where fish are caught. With this record in hand, temporal comparisons of the
human samples show a more intensive focus on freshwater fish in the Early Neolithic
(8000–6800 yBP) compared to a wider dietary breadth in the Late Neolithic and Early Bronze
Age (6200–4000 yBP).
The record of dietary reconstruction via bivariate plots has clearly provided an invaluable
reference point for interpreting past diets. Bivariate plots of δ13C from collagen and carbonate
have also been established, based on controlled feeding experiments (Kellner & Schoeninger,
2007). Regression models derived from these experiments provide a powerful source for
developing considerable precision to dietary reconstructions and serve to explain why a
bivariate characterization of isotopes is important. That is, collagen and carbonate captures
the respective fish and millet in the diet, but the model captures both.

8.2.5 Isotope chemistry: inferring patterns of residential mobility


The foregoing discussion makes clear that some key attributes of stable isotopes could be used
to document place of residence and evidence of residential change and migration. Variation in
both carbon and nitrogen isotope composition is highly influenced by habitat, and the same is
true for at least two other isotopes, strontium and oxygen, the former relating to local geology
and the latter to climate and especially water source (Bentley, 2006; Katzenberg, 2008; Price
et al., 2002). The study of residence, residential change, and migration have a long history in
anthropology, with regard to both living and past populations, in large part because of the
strong interest in understanding the origins and evolution of particular behavioral, cultural,
technological, and social variation in the history of our species. That is, understanding
residential change and patterns of mobility provide essential context for understanding key
developments in specific regions of the world. If it is possible to distinguish migrants from
local individuals in a population based on the isotopic signatures, then that becomes a record
for documenting and interpreting sources of variation for a range of issues, including

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 339

especially migration and mobility (Price, 2008). Identification of non-local individuals in a


population also has implications for understanding demographic characteristics, prevalence
of pathological conditions relating to infectious disease, and identification of social classes.

Strontium isotopes
Strontium has proven especially robust in identifying individuals who may be migrants and
for identifying migration during a person’s lifetime by comparing isotopic variation of earlier
forming tissues with later forming tissues, such as teeth and bone in adults or earlier forming
teeth and later forming teeth (or bone) in juveniles. Bones and teeth are both used, but owing
to its lower potential for diagenesis, tooth enamel is the preferred tissue for analysis.
Strontium is present naturally in the geological environment, including four naturally occur-
ring isotopes – three stable (84Sr, 86Sr, and 88Sr) and one radiogenic, unstable (87Sr) isotope
that is formed by the decay of 87Rb. Strontium enters the ecological system, originating in
bedrock and then moving into soil and groundwater, and subsequently enters the food chain
(Price et al., 2002). However, unlike the other elemental isotopes described earlier, strontium
isotopes have very small relative mass differences – the isotope composition remains rela-
tively unchanged from its origin through the food chain (but see Knudson et al., 2010). Thus,
unlike carbon and nitrogen stable isotope ratios, there is not a diet-to-tissue spacing seen, for
example, in trophic level changes in nitrogen. The strontium isotope ratios determined from
archaeological skeletal and dental tissues directly reflect the local geochemistry, and the
geochemical isotope ratio is passed unaltered from food to consumer (Ericson, 1985; Price
et al., 2002; Sealy et al., 1991).
Geochemistry of the decay process has revealed that the ratio of 87Rb/87Sr is directly
proportional to the ratio of 87Sr/86Sr, and it shows considerable variation in local geology
in some regions, with earlier forming rocks having higher ratios of 87Sr/86Sr than in later
forming rocks (Bentley, 2006). When local and regional geochemistry show distinctive
patterns of values in local settings and regions, the ratio of 87Rb/87Sr provides an invaluable
tool to identify migrants in a population. For example, owing to the common geological
history in Denmark and across southern Scandinavia, northern Germany, and northern
Poland, the 87Sr/86Sr values are generally homogeneous with values ranging from about
0.7090 to 0.7108 (Price et al., 2011).
The values of human tissues derived from archaeological contexts are compared with the
local geological values. However, in practice, values derived from analysis of bones and teeth
are compared to bioavailable values of the burial context, including soil, archaeological
fauna, modern animals, and archaeological bones and teeth (Price et al., 2002, 2011). The
immediate analytical goal is to characterize the local population isotopically and to identify
individuals who are outliers (“foreigners”) from the local isotopic composition. Presumably,
these outliers represent persons originating from another provenience and having matured in
another geographic setting. These individual non-locals are identified as those whose stron-
tium isotope values fall outside the local range of strontium isotope variation (Bentley, 2006),
defined by some as plus or minus two standard deviations from the mean of modern, non-
ranging fauna isotopic ratios (Knudson, 2008) (Figure 8.19).
Initial bioarchaeological applications of these biogeochemical approaches to reconstructing
and interpreting residence in a handful of settings (Ericson, 1985, 1989; Grupe, 1995; Price,
Grupe et al., 1994; Price, Johnson et al., 1994) revealed that variation in strontium isotope

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
340 Isotopic and elemental signatures of diet, nutrition, and life history

0.720
Tiwanaku Tilata, Coya Coya-3 Solcor-3 Chen Chen
0.719 Kirawi, Oriental
Iwawe
0.718

0.717

0.716

0.715

0.714
87Sr/ 86Sr

0.713

0.712

0.711

87
0.710 Sr/ 86Sr = 0.7087−0.7106 for
southeastern Lake Titicaca Basin
0.709

0.708

0.707

0.706

0.705

Figure 8.19 Strontium isotope ratios (87Sr/86Sr) from human tooth enamel samples across
Tiwanaku sites, South Central Andes. The horizontal gray bar captures the local range in
strontium isotope ratios for Lake Titicaca Basin, which is determined by sampling strontium
isotope ratios from local archaeological fauna. Isotopic values from human teeth enamel,
which fall within this range (e.g., seven individuals from the Tiwanaku sample), indicate
these individuals lived within the Lake Titicaca Basin during childhood. (From Knudson,
2008.)

ratios can inform our understanding of population dynamics as they relate to identification of
non-locals and locals in archaeological skeletal series. A pilot study of isotope ratios of dental
enamel and bone from late prehistoric Grasshopper and Walnut Creek sites, Arizona shows
that only some individuals share values similar to local geology (Price, Johnson et al., 1994).
These individuals likely spent their lives at their birth residence, whereas others appear to have
been born elsewhere and moved to their place of residence at some later time. Comparison of
strontium ratios in tooth enamel and bone from the Bell Beaker period (2500–2000 BC) of
southern Bavaria also reveals significant variation (Price, Grupe et al., 1994), which appears to
be more pronounced in the earlier than the later part of the period (Grupe, 1995). This finding
suggests that populations in the later period may have been more sedentary, which is
consistent with settlement analysis from conventional archaeological data.
From both New World and Old World settings, there is a rich body of work showing
important elements of variation in strontium isotopes. Analysis of this variation reveals key
interactions ranging from intercommunity relationships to long-distance migration involving
relatively few individuals, such as in trading, or recruitment of individuals into some aspect of
society, such as the formation of military groups. In summary, identification of non-local
individuals via their bone chemistry in general, but especially via strontium stable isotope
ratios, provides a powerful means of illuminating social, political, and economic dynamics
locally and regionally.
In the Americas, there has been a long-standing interest in the influence of dominant,
complex societies on a region. In the South American Andes, during the last two millennia

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 341

prior to European conquest, a number of powerful state systems rose and dominated all or
parts of the modern countries of Peru, Ecuador, Bolivia, and Chile. In most of these settings,
the record of the rising state is well documented archaeologically by ceramic styles, iconog-
raphy, architecture, settlement pattern, and other material evidence. As the state develops, it
first originates in a “heartland” at a particular locality or set of localities representing various
communities. Over time, the state influences surrounding regions, sometimes covering many
thousands of square miles, including part or all of the Andes Mountains and coast. Anthro-
pologists and other social scientists are keenly interested in the mechanics of the rise and
spread of these early South American state systems. Did these states begin to dominate
surrounding societies via migration, the spread of ideology, violence and subjugation (see
Chapter 4), trade and commerce, or by some other means that is simply invisible in the
archaeological record? For example, in the Middle Horizon period of the South Central Andes
(AD 500–1000), a very detailed record shows that societies throughout the region were
influenced by the heartland of the Tiwanaku state, centered at the site of Tiwanaku near Lake
Titicaca, Bolivia. A number of archaeologists argued that this influence resulted from state
expansion and incorporation of a remarkably large region of the Peruvian Andes by the
Tiwanaku polity. By gradually colonizing the region, the state played a role in spreading
material culture and political, cultural, and economic systems (Kolata, 1993). Other authorities
propose that the conquest of the region was accomplished by commercial trade (Dillehay &
Núñez, 1988) and spread of ideology via what Browman (1978:327) referred to as “proselyt-
izing merchant missionaries.”
The general hypothesis that Tiwanaku and other state systems in the Andes in prehistory
involved colonization can be tested by documenting whether “foreign” individuals were
present in these communities via analysis of isotope ratios in the skeletons of the people
who were once members of the communities comprising the Tiwanaku state system. If it can
be demonstrated that non-local isotope ratios are present in a population displaying a
particular pattern of variation of local isotope ratios, then a case can be made for the presence
of foreign persons. However, that alone does not tell the entire story. Rather, it is the context
provided by other data that present the case for mechanisms of state development, rise, and
domination of a larger geography that included once-independent societies.
Analysis of strontium isotope ratios from sites representing communities having Tiwanaku-
style material culture and architecture in the Moquegua Valley of southern Peru – Omo, Río
Muerto, and Chen Chen – provides important documentation of the potential role of actual
movement of people (Knudson & Price, 2007; Knudson et al., 2004). In these communities,
architecture and material culture provide evidence to suggest that individuals living there
were either affiliated with or were originally from the Tiwanaku heartland. Omo, for example,
has the only Tiwanaku-style temple outside of the Tiwanaku heartland (Goldstein, 1992).
While the Tiwanaku signature in material culture – from pots to buildings – is ubiquitous
throughout the region, it remains possible that local people exclusively adopted it as they
came under the influence of the Tiwanaku state in the absence or very marginal presence of
individuals from the heartland. Analysis of strontium isotopes in dental enamel of individuals
from Chen Chen, Moquegua Valley (southern Peru) and Tiwanaku (Bolivia) reveal several
findings that suggest the movement of actual people. The 87Sr/86Sr baseline biogeochemical
ratios are distinctively different on comparing the two regions. That is, the 87Sr/86Sr ratios for
the Moquegua Valley range from 0.7059 to 0.7067, and the 87Sr/86Sr ratios for Tiwanaku,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
342 Isotopic and elemental signatures of diet, nutrition, and life history

Lake Titicaca basin range from 0.7083 to 0.7111. Analysis of strontium ratios from eight
individuals from Chen Chen reveals ratios that fall well outside the range for the Moquegua
Valley (Knudson & Price, 2007; Knudson et al., 2004). Two of the ratios, in fact, match the
isotope ratios that would be expected if they were from Tiwanaku. Both are women, and
the data suggest that when they were three to four years of age they were residents of the
Tiwanaku heartland. At the site of Tiwanaku, based on strontium isotope ratios, three of
the 10 individuals were non-local. Interestingly, all three were interred in an elite quarter of
the monumental district of Tiwanaku. They included a young adult female and a young adult
male, both likely sacrificial offerings for the dedication of important ritual structures. The
third individual, also an adult, was a dedicatory offering at another ritual structure.
Weathering on the skeleton suggests that this body may have been displayed publically prior
to interment (Blom et al., 2003; Knudson & Price, 2007; Knudson et al., 2004).
The isotopic record from Chen Chen and Tiwanaku reveals that non-local individuals are
present at both centers. While the data sets are small, they nevertheless imply the presence of a
larger network involving movement of people from a center (Tiwanaku), taking place after the
formation of teeth. Regardless of scope, these analyses support the hypothesis that non-local
individuals are present at two important communities in the larger sphere of the
Tiwanaku state.
This is not to say that all of Tiwanaku territory – heartland and hinterland – involved actual
movement of people as a means of spreading political, social, and economic influence from
core to periphery. In this regard, communities from the San Pedro de Atacama oases and the
Loa River valley of northern Chile, while within the Tiwanaku sphere of influence, have no
individuals expressing non-local stable isotope ratios (Knudson, 2008; Knudson & Torres-
Rouff, 2009). Thus, in consideration of the Tiwanaku polity as a whole, while the influence of
the polity clearly shows that some communities were partially inhabited by individuals
originating from the core Lake Titicaca region and migrating to the periphery (e.g., the
Moquegua Valley), in other settings, local communities adopted elements of the dominant
society via other means (Knudson, 2008).
Isotopic variation as it relates to strontium documents the presence of non-locals within
communities and evidence of movement, likely explaining, in part, mechanisms of cultural,
social, and economic influence involving a dominant core polity. Similarly, an isotope record
of regional mobility documents the widespread influence of a dominant polity and long-
distance interaction in the Andes, including in states, such as the Wari empire (Conlee et al.,
2009; Knudson & Tung, 2011; Slovak et al., 2009; Tung & Knudson, 2008) and the Inca empire
(Turner et al., 2009). Other settings in pre-Hispanic Latin America also reveal a record of
movement of people, explaining, at least in part, the spread of a dominant system over
remarkable distances. At Teotihuacan (AD 1–650) in the Basin of Mexico near Mexico City,
archaeologists have identified foreign architecture centered in two ethnic compounds. The
skeletal remains are assumed to be foreigners living in these compounds (Price et al., 2000).
A relatively high degree of variation in stable isotope values in the tooth enamel but little
variation in the values in bone suggests that some individuals were born elsewhere and
migrated to the city in later life. This finding is consistent with the notion that the fast growth
of the city, while perhaps due, at least in part, to elevated fertility and birthrate, must also
have been fueled by significant migration (Price et al., 2000; and see Blanton, 1981). In at
least one setting in the city, numerous individuals display evidence of cutmarks and

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 343

mutilation patterns consistent with sacrifice. For example, at the Pyramid of the Moon, only
three of the 40 victims have strontium isotope ratios that suggest local derivation; all others
have values from a range of settings throughout Mesoamerica (Price et al., 2007).
While much of this record emphasizes aspects of empire building and social dynamics
under a variety of circumstances, it is also the case that inter-regional trade between cultures
in prehispanic Latin America is an important element in both of the settings described from
highly complex societies (Price et al., 2000; Wright et al., 2010). Similarly, it is important in
simple societies that are often incorrectly thought of as being isolated, such as island inhabit-
ants. However, stable isotope analysis reveals significant mobility and exchange in these
settings (Hoogland et al., 2010). Moreover, the strontium isotope record provides a means of
tracing the origins of enslaved individuals and patterns of forced migration of persons in the
African diaspora of postcolonial North America (Goodman et al., 2009; Price et al., 2006).
Investigation of stable isotope variation provides a window onto the role of migration in
the formation of new societies in the Old World, such as low mobility Holocene hunter-
gatherers in the southern Sahara Desert (Stojanowski & Knudson, 2011), late prehistoric
southeastern Arabia (Gregoricka, 2013a), Neolithic and Copper Age eastern Europe (Giblin,
2009), and Iron Age northeast Thailand (Cox et al., 2011). In the Oman Peninsula during the
third millennium BC, an abundance of foreign grave goods in local communal tombs led to
speculation that non-locals were frequently included as tomb members, an assumption
corroborated by the region’s growing involvement in inter-regional trade across the Persian
Gulf with major powers, including Mesopotamia and the Indus Valley. However, radiogenic
strontium isotope analysis revealed a different pattern, in which the majority of those interred
in these Bronze Age cairns were indeed local and displayed little isotopic variability indicative
of an increasingly sedentary lifestyle with the adoption of date palm agriculture (Gregoricka,
2013b) (Figure 8.20). Nevertheless, the few non-locals identified were buried according to
local customs, suggesting a form of “fictive kinship” as a means of strengthening political or
economic relationships (Gregoricka, 2013b).
This record contrasts with other settings showing relatively high levels of residential
mobility and population movement, colonization, and trade, such as in northwestern Africa
and the Near East (Buzon et al., 2007; Perry et al., 2008, 2009, 2012), and Imperial Rome
(Killgrove, 2010). Folklore, historical documentation, sagas, and archaeological documenta-
tion record the famously mobile Vikings of Scandinavia, who moved throughout northern
Europe, with travel extending throughout the continent and the northern Atlantic, and
including colonization of the British Isles, Iceland, Greenland, and North America, all by
the turn of the eleventh century AD (Fitzhugh & Ward, 2000; Jones, 1984). While widespread
movement was the case, identification of locals and non-locals in a number of settings reveals
alternative strategies of colonization and domination of regions that came under the influence
of the Vikings. In northwest Scotland in the New Hebrides islands, an area colonized by
Norwegian Vikings, four individuals in a sample of 20 were clearly distinctive in their isotope
ratios, enough to be identified as non-local (Montgomery & Evans, 2006). Although non-
local, their isotope ratios are strongly dissimilar to those from Norway, but may source these
individuals to either further south in England or Denmark. In contrast, a series of remains
from Viking-period Dublin reveal that all but one of the isotope values are local (Knudson,
O’Donnabhain et al., 2012). All values – from both enamel and bone – are remarkably
homogeneous and suggest that none of the individuals are from Viking populations

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
344 Isotopic and elemental signatures of diet, nutrition, and life history

0.7210

Umm an-Nar Island V


TA 161
Umm an-Nar Island I
0.7205
Umm an-Nar Island II
Unar I
Mowaihat B
0.7200
Tell Abraq
87Sr/ 86Sr

0.7195

0.7190

0.7185 MW 197

TA 161
0.7180
2700 BC 2500 BC 2400 BC 2300 BC 2200 BC 2000 BC

Figure 8.20 Strontium isotope ratios from human tooth enamel are presented for Bronze Age
tombs in Arabia. The local strontium isotope range, represented by the horizontal line with
standard deviations, was established by local faunal references. As the strontium ratios for
tooth enamel depict, individuals associated with these tombs spent their childhood years
within this local region. (From Gregoricka, 2013b; reproduced with permission of author
and Elsevier Ltd.)

originating elsewhere, such as Scotland or Scandinavia. These findings suggest an


acculturation model for Viking-period Dublin whereby social and cultural changes occurred
locally, rather than through immigration of a large number of first-generation individuals
from other areas of northern Europe (Knudson, O’Donnabhain et al., 2012).
The rise in the tenth century AD of the earliest kingdom in late Viking-period Scandinavia
involved the centralization of authority, a development that was facilitated by the construc-
tion of fortresses and other military outposts containing organized warriors. The first such
authority, King Harald Bluetooth, built fortresses and recruited males to station these fort-
resses in a number of settings in the Danish kingdom for control and administration of
territories. The “Trelleborg” fortresses have a distinctive and unusual circular structure so
different from regional architecture that they likely reflect a strong foreign influence. In order
to test the hypothesis that military and other residents of the fortresses were non-local, Price
and coworkers (2011) documented strontium isotope values in tooth enamel from skeletal
remains recovered from the Trelleborg fortress on the island of Zealand, Denmark. This
analysis included individuals from regular graves and individuals from a mass grave likely
resulting from deaths due to disease or accidents. Two-thirds of the 48 individuals revealed
values deriving from non-local contexts, certainly outside of southern Scandinavia
(Figure 8.21). At least four of the values are above the range and likely to come from Norway,
central and northern Sweden, or perhaps the northern British Isles (Price et al., 2011). These
findings are consistent with the hypothesis of mostly foreign-born individuals comprising

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.2 Isotopic analysis 345

0.720

0.716
87Sr/ 86Sr

0.712

0.708
Figure 8.21 Strontium isotope ratios from human (black bars) and faunal (gray bars, far left)
remains from Medieval Trelleborg, Denmark. Over two-thirds of the human isotope values
are outside the local geological strontium isotope range (horizontal gray bar), which
suggests that the majority of these individuals grew up on foreign, non-Scandinavian soil.
(From Price et al., 2011; reproduced with permission of authors and Antiquity.)

Harald Bluetooth’s military stronghold. In contrast to the Viking community at Dublin, the
Danish record as it is revealed in a military context is consistent with the high degree of
residential mobility in Viking-period Scandinavia – the military were recruited from a wide
variety of settings and there was considerable movement of individuals, sometimes involving
all ages and both sexes and with different health profiles in comparison with local and non-
local persons (Groves et al., 2013).
Interestingly, this pattern of widespread recruitment contrasts with the Imperial Roman
period pattern in the northern frontier British Isles where isotopic signatures indicate that
military recruiting was done on a local level, certainly mostly within the British Isles (Chenery
et al., 2011; regarding the diversity of the Roman diaspora in Britain, see Eckardt et al., 2009;
Evans et al., 2006; Hakenbeck et al., 2010; Leach et al., 2009; Montgomery et al., 2011;
Pollard et al., 2011).
Much of this record of individual movement and identification of foreign individuals
documents patterns of variation that speak to the movement of persons within the context
of wider population movements. Analysis of sex differences in adults (tooth enamel) provides
insights into marital residence patterns in past societies. In this regard, during the period of
agricultural intensification in Ban Chiang, Thailand, there is a strontium signature showing
considerably more variation in males than females (Bentley et al., 2005). This suggests a
matrilocal residence pattern where women grow up in the community consuming the same
foods, whereas men are drawn from a wider geographic spectrum and consumed different
foods. In contrast, females from Neolithic Central Europe have considerably more variation in
their strontium values than males, suggesting a patrilocal residence pattern (Bentley et al.,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
346 Isotopic and elemental signatures of diet, nutrition, and life history

2002). More precise information on biological patterns relating to marriage patterns and
inferences of residence are provided by biodistance analysis of genetic variation (see
Chapter 9). However, these highly distinctive patterns in strontium are strongly suggestive. In
this regard, at Ban Chiang, the strontium signatures may simply reflect differences in female
and male diets where girls and women restrict their food consumption to local sources and boys
and men have access to foods hunted some distance away. The combined records of strontium
isotopic signatures with biodistance analysis will help test these important hypotheses, espe-
cially with respect to processes behind the introduction and intensification of farming, which
must have been influenced by social structure and marriage residence patterning.

Oxygen isotopes
Climate has a tremendous influence on human adaptation, impacting such factors as resource
productivity and where people live. Thus, climatic patterns are potentially useful for inter-
preting land use patterns. δ18O values (determined from stable isotopes, 18O and 16O) of
terrestrial water vary in relation to climate, especially by temperature and humidity (Kolodny
et al., 1983; Luz et al., 1990). Because mammalian skeletal and dental phosphate and
carbonate are in equilibrium with body water, the phosphate oxygen isotope composition of
these hard tissues directly reflects the temperature and climate at the time of tissue formation
(Fricke et al., 1995). In order to test the hypothesis that hard tissues from archaeological
settings can be used to track the history of regional climates, Fricke and coworkers (1995)
determined temporal change in values from archaeological dental enamel (Eskimos and
Europeans) in coastal western Greenland and Denmark for the periods preceding, during,
and following the so-called “Little Ice Age” of the Medieval period (AD 1400–1700). There is a
3‰ decrease in δ18O values from earlier to later periods, which is consistent with increased
cooling documented in historical records describing climates in the North Atlantic
region. These changes are in accordance with other studies indicating increasing dietary
and climatological stress and eventual abandonment of Greenland by Vikings during the
Medieval period (Buckland et al., 1996; Scott et al., 1992).
There is a growing record to document residential mobility, place of origin (from enamel),
and later migration to a new geographic region (from bone or teeth that develop later in life),
thus making it possible to distinguish locals from non-locals (Katzenberg, 2008; White,
Spence et al., 2004). One of the neighborhoods at Teotihuacan on the western edge of the
city contains archaeological evidence – architecture and material culture – suggesting that it
was occupied by immigrants from Oaxaca. Consistent with the strontium isotope evidence
(Price et al., 2000), analysis of oxygen isotopes reveals values that document the clear
presence of non-locals, both in children (based on enamel) and in later life (based on bone),
including but not limited to immigration from Monte Albán in the Valley of Oaxaca (White
et al., 2007). Overall, the high degree of variation seen in the stable oxygen isotope values
expresses a picture of wide circulation of people in the region. Much of the relocation over the
course of lifetimes of individuals is due to normative behaviors related to trade, labor, and the
wide variety of behaviors responsible for movement in any society, past or present. In this
setting, analysis of oxygen isotope variation coupled with strontium isotope variation is
related to some very specific social behaviors. That is, analysis of one distinctive set of human
remains from Teotihuacan – those military or other guardians sacrificed and interred at key
ritual temples – provides important perspective on social events (White et al., 2007). While the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.3 Elemental analysis 347

motivation for the movement of individuals across the region to Teotihuacan is complex, at
least some of the movement of individuals is related to highly specific social practices,
including sacrifice.
As with strontium isotopes, analysis of oxygen isotopes facilitates the testing of long-
standing hypotheses about past social interaction, sometimes involving travel over consider-
able distances. While archaeological evidence based on material culture provides a snapshot
of regional interaction, in the absence of historical records, it does not provide the documen-
tation for mechanisms involving movement of people or the spread of ideas and influence in
the absence of migration. The strontium isotope record reveals important evidence for of
regions having complex social and population histories (see earlier). As with Mesoamerica, the
oxygen isotope record is consistent with the strontium isotope record showing variation in
immigration patterns, such as in Imperial Rome (Prowse et al., 2007) and in Roman colonies
such as the British Isles (Eckardt et al., 2009; Evans et al., 2006; Leach et al., 2009).

8.3 Elemental analysis


8.3.1 Background
Various elements found in both the organic (collagen) and inorganic (mineral or apatite)
components of bone tissue have dietary and nutritional significance. Most of these elements
are contained in the inorganic component (Burton, 2008; Sandford, 1992); more than 99% of
strontium in vertebrates, for example, is found in the bone mineral (Schroeder et al., 1972).
Major or bulk elements found in both the organic and inorganic components (carbon,
hydrogen, iron, nitrogen, calcium, phosphorus, oxygen, potassium, sulfur, chlorine, sodium,
and magnesium) perform critical functions in structural maintenance and are generally
required in relatively large quantities by humans. Trace elements serve mostly in catalytic
reactions and are frequently associated with certain enzymes, such as metal-activated
enzymes. Zinc, manganese, and iron serve vital functions. Some trace elements are toxic, if
ingested at even very low levels (e.g., arsenic, lead, mercury, cadmium), although virtually all
trace elements are toxic if taken excessively.
The biochemistry and physiology of many major and trace elements in humans are well
known, especially in relation to specific dietary properties and insufficiencies. A series of
pioneering studies established the importance of trace elements for identifying and interpret-
ing diet in past populations (Brown, 1973; Gilbert, 1975; Schoeninger, 1979). Initially, this
research was received with considerable enthusiasm, especially because it was assumed that
trace element values represent accurate and unaltered signatures of past diets, which had not
been identified by any other means in archaeological settings. Accumulating evidence shows,
however, that trace elements and their interpretation in archaeological human skeletal
remains are far more complex than had originally been anticipated. The early enthusiasm
for elemental analysis has been tempered by the realization that a number of factors, such as
food preparation techniques, cooking utensils, geochemical variation, physiological processes,
synergism between elements, inter- and intra-bone variation, age, sex, and especially dia-
genetic processes or postmortem alterations following death can alter the chemistry of bone
tissue in profound ways. Thus, trace elemental analysis research is devoted to distinguishing
between diagenetic and biogenic signals in archaeological remains (Burton, 2008; Lambert

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
348 Isotopic and elemental signatures of diet, nutrition, and life history

et al., 1990; Pate et al., 1989; Price et al., 1992; Sandford, 1993b; Sillen, 1989). Analyses
generally focus on three areas: (1) alkaline-earth elements (strontium and barium); (2) multi-
element analysis; and (3) single-element analysis.

8.3.2 Alkaline-earth elements: strontium and barium analysis


Strontium
Unlike the isotopic composition of strontium, which does not fractionate by biological (or
geological) processes, elemental concentrations show broad variation by trophic position in
plants and animals (Burton, 2008; Price, Grupe et al., 1994). Strontium is taken up by
organisms in inverse proportion to their trophic level. Among numerous elements utilized
in paleodietary and bioarchaeological research, strontium is “the only firmly established
elemental model in bone-chemistry analysis” (Ezzo, 1994a:608). Strontium has no known
biochemical function. It resembles calcium (also an alkaline-earth element) structurally and
can substitute for calcium in a number of physiological roles, including fixation in the
hydroxyapatite crystal structure of calcified tissues (Likins et al., 1960; Rosenthal, 1981). At
the bottom of the food chain, plants acquire strontium directly through soil, whereas
mammals, including humans, obtain the element through secondary sources, such as plants
or the animals that consume plants. Additionally, mammals discriminate against strontium in
favor of calcium. This discrimination pattern indicates that mammalian tissues contain less
strontium than plants, herbivores contain less strontium than plants, carnivores contain less
strontium than herbivores, and omnivores (such as humans) are intermediate between herbi-
vores and carnivores. Trophic level distributions of strontium have been observed in field
studies, both aquatic and terrestrial (Elias et al., 1982; Ophel, 1963; Schoeninger, 1985).
Various local factors influence strontium chemistry, including geology and regional levels
of alkaline-earth elements, and there is a considerable amount of variability in strontium (and
Sr/Ca ratios) within trophic levels (plants, herbivores, carnivores) (Burton, 2008; Runia,
1987a; Schoeninger, 1985; Sillen, 1992; Sillen & Kavanagh, 1982; Sillen & Smith, 1984).
Even within the same environment, economically important cereals and roots may have either
elevated or lowered Sr/Ca ratios in comparison with other plants (Runia, 1987a). Owing to
variation within trophic levels, knowledge of local foodwebs and predator–prey relations is
essential for accurate reconstruction of past diets based on elemental strontium (Burton,
2008). In addition, because bone strontium and Sr/Ca ratios can be influenced by consump-
tion of high-calcium foods (e.g., seafood) and use of mineral additives, the simple measure-
ment of strontium in archaeological bone is not necessarily synonymous with trophic level
and relative proportion of plant to meat consumption in past populations (Burton, 1996, 2008;
Burton & Wright, 1995; Ezzo et al., 1995). For a variety of settings, strontium may not be a
very accurate indicator of trophic level or meat versus plant consumption. Owing to its
chemical bonding properties with calcium, it represents the dietary sources of calcium
(Burton, 1996, 2008). This and other confounding factors indicate that inter- and intra-
population comparisons based on archaeological samples must be made with a great deal of
caution (Buikstra et al., 1989; Burton, 1996, 2008; Ezzo, 1994a; Radosevich, 1993; Sandford,
1992, 1993b; Sillen & Kavanagh, 1982; Sillen et al., 1989).
Researchers have also become increasingly aware that diagenesis in elemental analysis is
the most important impediment to paleodietary reconstruction (Ezzo, 1992; Grupe &

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.3 Elemental analysis 349

Piepenbrink, 1988; Lambert et al., 1983, 1984; Nelson & Sauer, 1984; Nelson et al., 1986; Pate
& Hutton, 1988; Pate et al., 1991; Price, 1989; Sillen, 1981, 1992; and others). In addition to
the aforementioned factors influencing strontium in bone, there are a number of behavioral
and dietary influences. A variety of evidence shows that weaning (Sillen & Smith, 1984),
pregnancy and lactation (Blakely, 1989; Radosevich, 1989, 1993), consumption of some
nondomesticated plants (e.g., nuts) and shellfish (Benfer, 1984; Buikstra et al., 1989; Byrne
& Parris, 1987; Kyle, 1986; Schoeninger & Peebles, 1981), and health (Dolphin et al., 2005;
Lane & Peach, 1997) will influence values. In summary, reliable biological signals can only be
acquired through careful consideration of a range of factors influencing strontium in bone
tissue. Much of the literature documenting strontium in archaeological bone has not con-
trolled for diagenesis and other factors. Thus, the results based only on strontium should be
viewed with skepticism (Katzenberg, 1993).
Studies of strontium provide results that are suggestive of dietary strategies (Katzenberg,
1984, 1993; Radosevich, 1989, 1993; Runia, 1987b; Schoeninger, 1981, 1982; Sillen, 1981;
and others). Using strontium values from herbivore and carnivore controls, comparisons of
Aurignacian foragers with later Natufian farmers at Hayonim Cave, Israel, show dietary
change consistent with the shift to agriculture (Sillen, 1981). Analysis of other Middle
Eastern populations (Schoeninger, 1981) and comparisons of archaic and modern humans
(Schoeninger, 1982) also demonstrate marked distinctions in dietary focus between foragers
and farmers. Both the Sillen and Schoeninger investigations demonstrate the necessity of
using herbivore and carnivore control samples when attempting to characterize human diets
in local settings. Through her comparison of human and herbivore strontium values, Schoe-
ninger concluded that “[w]ithin the Levant, the nonagricultural populations. . .were including
a large proportion of plant material in their diets. This represents increased use in plant
material when compared to the earlier human population [in the region]” (1981:87). This
conclusion was independently confirmed using a similar research protocol based on analysis
of strontium/calcium ratios from humans, herbivores, and carnivores at Hayonim Cave
(Sillen, 1981; Smith, Bar-Yosef et al., 1984).
In prehistoric Ontario populations studied by Katzenberg (1993), where diagenesis and
other factors influencing bone strontium are controlled for, there is a decrease in values that
reflects a shift from plants high in strontium (leafy plants and nuts) to low-strontium grasses
(i.e., maize) (Figure 8.22). The results are consistent with findings based on nitrogen and
carbon stable isotopes (see earlier), although the approaches – isotopic and elemental – refer to
different parts of diet, those that contribute to the organic and inorganic fractions of bone
tissue.
Schoeninger (1979) determined strontium values in members of different status groups at
Chalcatzingo, Mexico. Her study suggested lower values in higher-ranked individuals, which
is consistent with the hypothesis that the elite in this society likely had greater access to
animal protein whereas lower-ranked individuals consumed relatively more plants, such as
maize (Blakely & Beck, 1981; Hatch & Geidel, 1985; Jacobs, 1995, for alternative patterns).
Weaning patterns have also been estimated on the basis of Sr/Ca ratios by mapping the
variation in values from earliest to latest forming enamel. Spatial analysis of Sr/Ca ratios from
deciduous teeth in healthy European children of known dietary/nursing histories shows clear
patterns of change (Humphrey et al., 2008). The pattern is highly specific, documenting the
onset of breastfeeding at birth (decrease in Sr/Ca ratios), its duration, and the introduction of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
350 Isotopic and elemental signatures of diet, nutrition, and life history

250

200
Strontium (‰)

150

100

50
0 500 1000 1500 2000
Year (AD)
Figure 8.22 Plot of trace element (strontium) values in prehistoric and historic Ontario
native populations, showing an increase in maize in diets after AD 1000. (Data from
Katzenberg, 1984.)

solid foods (increase in Sr/Ca ratios). While not applied to archaeological remains, this
approach has considerable promise for tracking life history as it relates to diet.

Barium
Barium has received little attention in bioarchaeological chemistry and paleodietary studies.
This underrepresentation is surprising as a significant body of evidence indicates that the
element is a highly sensitive indicator of past foodways (Burton & Price, 1990a, 1990b; Ezzo,
1992, 1993, 1994a; Francalacci & Borgonini Tarli, 1988). Like strontium, barium has no
known biochemical function, it is nontoxic, it is structurally similar to calcium, it is not
tightly regulated metabolically, and it undergoes biopurification with increasing trophic level
(Burton & Price, 1990a, 1990b; Burton & Wright, 1995; Elias et al., 1982). Barium values and
Ba/Ca ratios are lower in herbivores in relation to the plants they consume, and carnivores
have lower values and ratios than herbivores. Most barium in the geological environment
comes in the form of barite (BaSO4), a chemical that is relatively less soluble than the apatites
that are associated with strontium (Ezzo, 1992). This suggests that the soil-to-plant discrimin-
ation should be greater for barium than for strontium. Consistent with this hypothesis,
analysis of hundreds of archaeological bone samples indicates that barium separates organ-
isms into their trophic positions better than does strontium (Burton & Price, 1990a, 1990b).
Thus, barium may be an even more sensitive indicator of diet than strontium.
The importance of barium has been shown by the comparison of the analysis of Sr/Ca and
Ba/Ca ratios in animal and human bone from early prehistoric foragers from Carrier Mills,
Illinois. The ratio Sr/Ca indicates only slight differences between animal species; the overlap-
ping ranges of values in herbivores and carnivores prevent any meaningful dietary interpret-
ation (Burton & Price, 1990a). Barium ratios (Ba/Ca) show clear interspecies differences.
Human values at Carrier Mills are closer to values for carnivores than herbivores, suggesting
that meat is a dominant food in the diets of these prehistoric foragers.
Barium analyses yield an important perspective on temporal shifts in diet in relation to
social and environmental circumstances (Ezzo, 1992, 1993, 1994a, 1994b). Increased

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.3 Elemental analysis 351

population aggregation and environmental degradation (wood depletion, increased aridity,


game depletion) in later prehistory appears to have evoked a shift toward reduced dietary
diversity and decreased hunting over the course of the 125-year occupation of Grasshopper
Pueblo (Ezzo, 1992, 1993). Decreased barium values indicate an increased use of maize and
decreased use of meat in the diet, much more so in females (from 350.1 ppm to 261.1 ppm)
than in males (from 285.6 ppm to 271.8 ppm) (Ezzo, 1992, 1993). These findings indicate that,
in the earlier period, females consumed more wild plants and less maize and meat than did
males. This pattern of dietary sexual dimorphism is similar to dietary inferences made from
observations of pathological markers (e.g., dental caries). In the later period, female and male
diets became virtually identical, with both sexes placing increased emphasis on maize. Thus,
most of the dietary change appears to be associated with females. The convergence of female
and male diets in the later period may reflect increased involvement of men in agricultural
activities, a pattern of activity that characterizes agricultural intensification in various
ethnographically observed populations worldwide (Ember, 1983). The reasons for a shift
toward a greater focus on maize are speculative. It may be that maize was a more reliable
food resource in the face of increasing aridity and environmental stress during the later
occupation of the region (Ezzo, 1992, 1993).
Burton and Price (1990a, 1990b) reviewed a wide range of published data on the barium
content of rocks, soils, plants, fresh water, and sea water (Wessen et al., 1977, 1978). The
pattern that emerges from their review is one in which barium and strontium values are
approximately equal in terrestrial settings, but barium and Ba/Sr ratios in marine and
terrestrial settings are highly distinctive (Figure 8.23). Barium values and Ba/Sr ratios are
considerably lower in sea-water and marine organisms than in terrestrial organisms. Analy-
sis of human remains from a wide range of New World archaeological sites from predomin-
antly marine settings, coastal sites with agricultural (terrestrial) consumption, and inland
sites with no access to marine foods reveals that Ba/Sr ratios also clearly distinguish
between human populations utilizing marine versus terrestrial resources (Burton & Price,
1990a, 1990b). The ratios in archaeological human bone samples from desert settings (e.g.,
Stillwater Marsh, Nevada) are an exception to the terrestrial pattern, and appear to be more
similar to marine values. This unusual pattern may result from strontium enrichment in
desert soils, which would immobilize barium but not strontium (Burton & Price, 1990a,
1990b). Thus, in at least some desert settings, differentiation of marine/terrestrial diets based
on Ba/Sr ratios does not appear possible. All other previously tested contexts clearly separate
human populations that use predominantly marine resources from those using terrestrial
resources.
Like strontium, barium offers an opportunity to understand changes in diet in early life by
documenting spatial variation in tooth enamel. Analysis of Ba/Ca ratios of healthy children
and captive macaques reveals that, like Sr/Ca ratios (Humphrey et al., 2008), there is a clear
pattern of change beginning at birth (based on location of the neonatal line) and terminating
with the final formation of enamel (Austin et al., 2013). In particular, the Ba/Ca ratio increases
in the first three postnatal months (the period of reliance on breast milk). Following the peak
in Ba/Ca ratio values, there is a decrease in values reflecting the decline in and termination of
consumption of breast milk. The high degree of correlation of values is revealed in the slow
reduction of ratios in gradual weaning versus the rapid reduction in ratios for individuals
where weaning is terminated quickly.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
352 Isotopic and elemental signatures of diet, nutrition, and life history

0.25

0
12
7 11
–0.25 10
8 9
–0.50
Log (Ba/Sr)

–0.75 Marine diet Terrestrial diet

-1.0

-1.25 13 14
4
-1.50 6
1 3
5
-1.75 2

-2.0

Figure 8.23 Ba/Sr values in archaeological bone, showing the distinction between marine
and terrestrial consumers: (1) Paloma, Peru; (2) Rolling Bay, Alaska; (3) Kiavak, Alaska;
(4) Three Saints Bay, Alaska; (5) Chaluka, Alaska; (6) Port Möller, Alaska; (7) Rio Viejo,
Mexico (coastal); (8) Cerro de la Cruz, Mexico (coastal); (9) Fabrica San Jose, Mexico;
(10) Monte Albán, Mexico; (11) Poland (multiple sites); (12) Pirincay, Ecuador;
(13) Pueblo Grande, Arizona; (14) Stillwater Marsh, Nevada. (From Burton & Price,
1990b; reproduced with permission of authors and Academic Press Ltd.)

8.3.3 Multi-element analysis


Additional discrimination of dietary focus in past populations has been explored by analysis of
multiple elements, an approach first advocated by Gilbert’s study of human bone samples from
the Dickson Mounds site, Illinois (Ezzo, 1994a; Gilbert, 1975, 1977, 1985; and see Crist, 1995;
Sandford, 1992, 1993b; Smrčka, 2005 for discussions of other examples of multi-element
analysis in different settings). Among the five elements included in his analysis – copper,
strontium, magnesium, manganese, and zinc – the element that best discriminated between
nonintensive agriculturalists and intensive agriculturalists from the earlier and later occupa-
tions of the site, respectively, was zinc: the earlier group had higher values than the later group.
Lambert and coworkers (1979, 1984; and see Szpunar et al., 1978) employed a range of
elements in documenting the shift from foraging to maize farming in later prehistory in the
lower Illinois Valley. The detailed analysis of bone and soil samples from the burial matrix serves
to underscore the potential of diagenesis and elemental soil–bone and bone–soil transfer. In large
part, most elements show some evidence of diagenesis, thus not rendering insights into past diets.

8.3.4 Single-element analysis: tracking dietary deficiency and toxicity


Iron
Deficiency in iron in past populations is inferred via analysis of pathological lesions (cribra
orbitalia, porotic hyperostosis) in a range of settings worldwide (see Chapter 3). Some suggest
that elemental analysis of archaeological bone samples provides additional understanding of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.3 Elemental analysis 353

iron status where skeletal lesions have been documented (Edward & Benfer, 1993; Fornaciari
et al., 1981; Sandford et al., 1988; Zaino, 1968). Based on limited samples, the presence of
lower iron in individuals with cribra orbitalia in Punic Carthage (third century BC) compared
with nonpathological individuals suggests a link between iron status and pathological indi-
cators of anemia (Fornaciari et al., 1981; but see Richtsmeier & Sheridan, 1996).
Juveniles from later Carthage (seventh century AD) have significantly higher levels of iron
than adults, which may indicate that iron in archaeological bone represents a biogenic signal
(Sandford et al., 1988). Comparison of iron levels in the third century BC and seventh century
AD Carthaginians shows higher values in the later period. This suggests that either iron status
improved or iron intakes increased in the later period (Sandford et al., 1988).
Ezzo (1994a) has questioned the biological significance of iron measurement in archaeo-
logical human bone. Except with regard to marrow, iron has no apparent function in bone
tissue, nor does it act as a reservoir for the element (Ezzo, 1994a). The lack of physiological
evidence suggests that varying amounts of iron in archaeological bone likely do not correlate
with prevalence of iron deficiency anemia. Additionally, iron content in soil is considerably
higher than in bone, indicating that unless all dirt is removed – a virtual impossibility in
archaeological remains – iron content in the skeletal tissue is highly exaggerated (Carvalho
et al., 2004; Ezzo, 1994a).

Zinc
Following Gilbert’s (1977) lead, there has been a general acceptance of zinc as an important
discriminator of diet (e.g., amount of meat versus plant consumption) and inferences about
nutritional health in past populations. Unlike strontium, calcium, or barium, a theoretical
basis has not been developed for zinc as a paleodietary indicator (Ezzo, 1994a, 1994b). Thus,
although zinc may be sensitive to dietary history, the lack of theoretical models limits its
usefulness in bioarchaeological chemistry and dietary inference (Burton, 2008). Underlying
the use of zinc is the unsubstantiated assumption that the element is present in higher levels in
meat and shellfish than in plants. Zinc acquired in the diet will influence the levels in skeletal
tissues of growing animals, such that diets deficient in zinc will result in low values in bone
and other tissues (Ezzo, 1994b). Zinc appears to play an essential role in growth: severe
deficiencies of the element result in markedly reduced growth and dwarfism. Thus, in contrast
to strontium and barium, the essential nature of zinc and its rather tightly controlled
metabolic regulation suggest that it has little use as a paleodietary indicator.

Lead
Lead is among the most discussed elements in bioarchaeological chemistry, mostly in
attempting to document chronic toxicity and excess intake in past groups (Aufderheide,
1989; Aufderheide et al., 1981, 1985; Aufderheide, Wittmers et al., 1988; Ericson et al.,
1979, 1991; Jaworowski, 1968; Keenleyside et al., 1996; Lalich & Aufderheide, 1991; Mont-
gomery et al., 2010; Nakashima et al., 2011; Perry et al., 2012; Polet et al., 2000; Reinhard &
Ghazi, 1992; Reinhard et al., 1994; Smrčka, 2005; Waldron, 1981, 1982, 1983; Wittmers et al.,
2008). Like barium and strontium, lead shows preferential bone deposition in comparison with
other tissues (more than 95% of absorbed lead is stored in the skeleton), it is not readily
excreted, and it has an unusually low turnover rate in skeletal tissue (Aufderheide, 1989;
Montgomery et al., 2010; Moore, 1986; Sandford, 1992; Wittmers et al., 1988). These factors
indicate that adult skeletal lead values represent a measure of lifetime exposure. Because

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
354 Isotopic and elemental signatures of diet, nutrition, and life history

exposure is only rarely due to natural sources, lead found in bones and other tissues originates
from anthropogenic factors (Aufderheide, 1989; Burton, 2008; Montgomery et al., 2010;
Smrčka, 2005), but only if diagenetic sources can be ruled out (Wittmers et al., 2008).
In Colonial-era North America, the use of pewter – a lead-based material used in cooking
and eating implements – resulted in increased exposure to lead, in some instances at toxic
levels (Aufderheide et al., 1981, 1985; Aufderheide, Wittmers et al., 1988). Owing to greater
access to pewter by elite individuals in Colonial society, they had greater risk of exposure to
lead than did non-elite individuals. The relatively higher exposure is well illustrated in the
comparison of the bones of land owners and their slaves. For example, family members of the
owner of the Clifts Plantation, Virginia, had a considerably higher lead content than their
slaves (185 ppm vs. 35 ppm) (Aufderheide et al., 1981). This distinction reflects the use of
pewter by the landowner class and wood or unglazed ceramics by the non-elite class. Analysis
of lead content in military burials from the War of 1812 Snake Hill sample, Ontario, also show
low values suggesting that recruits were drawn from lower socioeconomic levels of the
civilian population (mean ¼ 31.3 ppm; Lalich & Aufderheide, 1991). Similarly, high-status
individuals from Medieval Belgium (c. AD 1100–1500) and Edo-period Japan (AD 1603–1867)
have elevated lead content, reflecting lead toxicity (Nakashima et al., 2011; Polet et al., 2000).
The former may be due to use of tin pewter eating utensils. The latter is exhibited as lead
bands in the metaphyses of femora in young juveniles. These indicate the incorporation of
lead in the growth plates during the developmental years. While the source is unclear, given
the very young age of these individuals (<3 years old), breastfeeding mothers may have
inadvertently poisoned infants and young children.
Slave skeletons from a Colonial-era Barbados sugar plantation have very high skeletal lead
concentrations (mean ¼ 118 ppm; values up to 424 ppm) (Handler et al., 1986, 1988). Skeletal
lead in this setting is apparently derived from various origins (e.g., pewter and lead-glazed
vessels used for storage, preparation, or serving of food and drink). The major source of lead
originated from the processing of sugar into rum via lead stills. In a similar vein, unusually high
intake of lead in Romano-British and later populations in Great Britain is well documented
(Montgomery et al., 2010; Waldron, 1981, 1983) (Figure 8.24). The source of lead is unknown,
but it appears to have been acquired from use of lead-based eating utensils, water pipes, and
simply living in a setting involving mining or metal-using, thereby increasing lead exposure.
During the late eighteenth and early nineteenth centuries, Omaha populations from north-
eastern Nebraska used lead for a variety of purposes – such as for the production of body
ornaments and musket balls (Reinhard & Ghazi, 1992; Reinhard et al., 1994). Analysis of
skeletal remains from this group reveals unusually high lead concentrations in some individ-
uals. Analysis of stable lead isotope ratios (206Pb/204Pb) suggests that lead was traded to the
Omaha from present-day Missouri. Lead was also used to make red facial pigment, which would
have been readily absorbed. Thus, facial paint is likely the primary source of lead for many of
these individuals. Some skeletal lead may have also originated from activities taking place after
death, since red paint was also applied to deceased individuals prior to their interment.

Mercury
Mercury is a highly toxic heavy metal that, when ingested either in its element (Hg) form or in
compound form (e.g., mercury sulphide), causes mercury poisoning (hydrargyria), a condition
resulting in neurological disruption and death. Mercury can be ingested, either through food

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
8.4 Methodological issues in bioarchaeological chemistry 355

Median concentration (mgKg t–1)


4
1st–3rd c. AD Rome

0
ic e e l
lit
h Ag Ag
e
AD AD AD va
Ag ie
eo nz
e
Iro
n n rie
s
rie
s
rie
s ed
N o Iro u u u M
Br sh nt nt nt te
/Ir i ce ce ce La
h h h
tish 4t 7t 1t
ot t– h– 1
1s 5t h–
Sc 8t
Period

Figure 8.24 Mean lead concentrations in human tooth enamel showing increased exposure
to anthropogenic lead from Neolithic to Late Medieval periods in Britain. The average
lead concentration in enamel from the first to third century AD Rome is depicted for
comparison. (From Montgomery et al., 2010; reproduced with permission of authors and
Cambridge University Press.)

or directly. Regarding the latter, during the Medieval period, mercury was prescribed for
treatment of venereal syphilis and leprosy. Atomic absorption analysis of a large series of
skeletons from southern Denmark having osteological evidence of syphilis and leprosy (see
Chapter 3) revealed that nearly half of the individuals likely having syphilis showed elevated
mercury levels (Rasmussen et al., 2008). These elevated levels were considerable and suggest
that in life these individuals had been treated with mercury. Similarly, nearly all individuals
having evidence of leprosy had mercury levels strongly suggesting that they were also treated
with mercury. With several exceptions, individuals lacking leprous or syphilitic lesions had
either no or minimal mercury levels. The exceptions pertain to three individuals recovered
from the friars’ cemetery of the Svenborg Friary who showed toxic levels of mercury.
Rasmussen and coworkers (2008) suggest that these individuals absorbed mercury, either
through the skin or via inhalation, while using mercury-laden red ink. Alternatively, these
individuals may have been exposed to mercury during the process of its administration to sick
individuals in the Abbey hospital.

8.4 Methodological issues in bioarchaeological chemistry


Both of the primary components of bone tissue, apatite and collagen, provide useful dietary
information. In comparison with isotopic analysis of collagen, elemental and isotopic analysis
of the carbonate in bone apatite is more problematic because of the stronger influence of
diagenesis and its myriad confounding factors in post-burial environments (Hedges, 2002).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
356 Isotopic and elemental signatures of diet, nutrition, and life history

Stable isotope (i.e., carbon and nitrogen) analysis of collagen has several advantages over
elemental and stable isotope analysis of apatite in dietary documentation and interpretation.
Importantly, because bone collagen is not subject to ionic exchange for carbon and nitrogen,
diagenetic effects are not as influential in determining values as in bone apatite (Ambrose,
1987; Grupe et al., 1989). Within the mineral component of bone, the apatite fraction is
subject to extensive diagenetic change (Schoeninger & DeNiro, 1982; Wright & Schwarcz,
1996). It is a more straightforward process to identify and remove contaminants (fats,
particulate plant matter, and humic matter) in collagen samples that potentially could alter
the biological signal (Ambrose, 1987, 1990; Ambrose & Norr, 1992; Stafford et al., 1988). New
technology and better understanding of the biochemistry of the mineral component (apatite)
of bone for trace element and stable isotope analysis has greatly expanded the range of
possibilities for paleodietary study (Metcalfe et al., 2009). Moreover, tooth enamel apatite is
relatively more resistant to diagenesis than bone apatite, and offers a source of potentially
valuable dietary information (Lee-Thorp, 2008). This is especially advantageous for humid
and/or low-latitude contexts (Krigbaum, 2007). However, there is no generally accepted
means of assessing diagenesis in tooth enamel (Grine et al., 2012; Schoeninger, 2010).

8.5 Summary and conclusions


The application of isotopic and elemental analyses facilitates an understanding of foodways
in past human societies by providing a nearly direct record of diet. This approach addresses
several issues in bioarchaeology and archaeology that have heretofore been essentially
speculative. Especially noteworthy are the timing and spread of C4 cultigens (e.g., maize
agriculture in the New World), the spread of agriculture generally (e.g., across Europe), the
relative contribution of marine and terrestrial foods in coastal settings globally, and the
highly localized nature of extraction of food sources. One of the most profoundly important
contributions of bone chemistry is the insight that analyses provide regarding migration and
social interaction within and across social and political boundaries. The identification of non-
locals from locals in populations via strontium and stable oxygen isotope variation has
provided the opportunity to revisit old or investigate new questions about a wide range of
regional population interactions in light of social, political, and culture history contextual
information. Both stable isotopes and trace elements are subject to sources of variation that
potentially impede dietary interpretation, including climate, habitat, physiology, and
diagenesis.
The approaches to dietary interpretations and ecological and behavioral inference discussed
in this chapter should be considered as part of larger issues dealing with subsistence, adapta-
tion, and nutritional ecology in the past. Although quantitative and qualitative means of
subsistence documentation (e.g., analysis of food refuse) are subject to a range of biases – as is
also the case for isotopic and elemental analyses – it is important to consider them in order to
gain a comprehensive perspective of the foodways of earlier societies. However, direct infor-
mation on diet at the individual level in particular and the impact of diet and nutrition during
the juvenile and adult years in general can be acquired only through the investigation of
human remains.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 07:09:43, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.010
Biological distance and historical
9 dimensions of skeletal variation

9.1 Introduction
Human remains are a central data source for documenting temporal and spatial patterning
of biological relatedness. Determination of biological relatedness has traditionally been
achieved via analysis of phenotypic skeletal and dental traits and their intra- and inter-
population variation. Biological distance or biodistance analysis is based on the premise
that the variation in morphological traits of the skeleton reflects underlying genotypic
variation. Although biodistance analysis pertains largely to characterizing phenotypic
variation, genotypic variation is becoming increasingly important via the study of ancient
DNA (aDNA).
Regardless of the type of biodistance data, the degree of relatedness presupposes that a
greater frequency of shared attributes indicates closer genetic affinity than a lower frequency
of shared attributes. In this regard, the orientation of the research is linked to the questions
being asked about intra- and inter-population interactions, ranging from continental-scale,
containing many sites, to regional-scale containing multiple sites, to single site analyses
focusing on individual variation within a population (Figure 9.1).
Biodistance analysis is motivated by three key interests (after Buikstra et al., 1990). First,
results are important for investigation of issues relating to evolutionary history, such as
genetic drift and selection, gene flow, and the influence of geography and other isolating
mechanisms on biological relatedness (Conner, 1990; Droessler, 1981; Heathcote, 1994;
Ossenberg, 1986; Rothhammer & Silva, 1990; Sciulli, 1990; Scott & Turner, 1997;
Stojanowski & Schillaci, 2006). By understanding the degree of relatedness, it becomes
possible to characterize temporal relationships in the identification of local in situ develop-
ment, or migration, or some combination thereof (Konigsberg, 2006).
Second, biodistance analysis addresses other fundamental archaeological and biohistorical
issues regarding the degree to which cultural and biological changes are influenced by
extrinsic factors versus local circumstances (Conner, 1990; Droessler, 1981). In such cases,
it is absolutely vital that population history is considered in order to fully assess the potential
contributions of these alternative agents. Biodistance analysis has the potential to identify
population boundaries, postmarital residence patterns, familial and kin groupings, social
groupings, and presence of individuals from other populations, especially in settings involv-
ing contact between biologically distinctive groups (Konigsberg & Buikstra, 1995) or long-
term regional gene flow (McIlvaine et al., 2014). Collectively, the biodistance record provides
important insights into population history, including shifts and stasis over time. Shifts imply
formation of new populations and social groupings (ethnogenesis) and stasis indicates lack of
local biological change with little or no external contribution of variation.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
358 Biological distance and historical dimensions of skeletal variation

(a) (b)

8.0 8
Plains Florida Panhandle
7
6.6 Illinois Florida North Interior

SW US 6 Florida South Gulf


Florida Florida Atlantic Coast
5.2

Variable 2
Variable 2

NW Coast Georgia Coast


4
3.8
Alaska
Georgia Interior
3
2.4
2

1.0 1
1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8
Variable 1 Variable 1

(c)
8.0
B. # 1

6.6 B. # 3

B. # 4
B. # 2
5.2
Variable 2

B. # 5
3.8

B. # 6
2.4

1.0
1 2 3 4 5 6 7 8
Variable 1

Figure 9.1 Biodistance analysis, employing the same two variables, on continental (a), regional (b), and
site (c) scales. (From Stojanowski & Schillaci, 2006; reproduced with permission of authors and John
Wiley & Sons, Inc.)

Third, biodistance investigations provide an important context for assessing skeletal and
dental variability where population structure may have an influence, such as disease or
nutrition (e.g., access to key nutrients and other resources by segments of a population; see
Buikstra, 1976a). In this regard, Buikstra noted that “It is well known. . .that the response to
both disease and generalized stress can vary from population to population. Therefore,
the strength of epidemiological statements must be tempered with a consideration of
biological relationships within and between the archaeological series” (1977b:72). The
adoption of new subsistence strategies by a population can have a tremendous influence
on the size of the population and where it lives. Resultant demographic shifts can, in turn,
influence patterns of mating and genetic exchange. In summary, the units of change – the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.1 Introduction 359

population groupings – identified through biodistance analysis provide an essential context


for bioarchaeological investigation.
Biodistance analysis from a broad variety of settings has served to identify patterns of
intra- and inter-population relationships, especially continuities across boundaries and over
time, and to provide context for understanding social and economic developments and
transitions (Alzualde et al., 2007; Bolnick & Smith, 2007; Carnese et al., 2010; Duncan,
2012; Forgey, 2011; Hubbe et al., 2009; Kemp et al., 2009; Konigsberg, 2006; Moraga et al.,
2005; Pilloud & Larsen, 2011; Sutter & Mertz, 2004; Ullinger et al., 2005; Varela et al., 2008;
and many others).
Biodistance analysis is complex, especially with regard to identifying meaningful patterns
of biological variation that distinguish individuals or groups, in either temporal succession or
spatial distribution. Because of the nature of pattern identification, considerable attention has
been devoted to methodological and theoretical concerns in determining biological affinity
(Irish, 2010; Konigsberg, 2006; Scott, 2008; Scott & Turner, 1997). In general, close biological
affinity is indicated by unusually high (or sometimes low) shared trait frequencies of specific
traits in some populations (Hrdlička, 1935; Jacobi, 2000; Larsen, Craig et al., 1995; Nelson,
1992; Snow, 1974) or within subgroups of populations (Alt & Vach, 1998; Pilloud & Larsen,
2011; Spence, 1994). Intra- and inter-population biological relationships are best identified by
simultaneous consideration of multiple traits via multivariate statistical analyses, which for
metric data include principal components analysis, canonical axes, or discriminant function
analysis, often coupled with subsequent cluster analysis or multidimensional scaling.
Although different biodistance statistics are used for metric and nonmetric (discrete trait)
data, these procedures have strategic commonalities. The multivariate Mahalanobis D2 statis-
tic has become a mainstay in the analysis of metric (continuous) data for population
inference, ethnicity, and temporal and spatial relatedness (Harris, 2008; Hemphill & Mallory,
2004; Irish & Konigsberg, 2007; Ross & Ubelaker, 2010; and many others). The mean measure
of divergence (MMD) statistic has also been used for much of the last half century for
nonmetric data, especially in dental applications from a wide variety of contexts focusing
largely on issues relating to migration and continuity within specific settings (Berry & Berry,
1967; Green & Suchey, 1976; Harris, 2008; Harris & Sjøvold, 2004; Irish, 2005, 2010; Lukacs &
Hemphill, 1991; Pietrusewsky, 1989; Sjøvold, 1973, 1977; Sutter & Verano, 2007; Torres-
Rouff et al., 2013; Ullinger et al., 2005) (Figure 9.2). The MMD statistic has been subject to
criticism, in part because it does not consider correlation between variables (Harris, 2008;
Konigsberg, 2006). However, it provides a realistic picture of relatedness and a generally
strong approach to biological affinity (Irish, 2010; Schillaci et al., 2009; Torres-Rouff et al.,
2013). Another powerful multivariate analysis uses principal component or coordinate scores
derived from tooth measurements for identifying broad-scale population relationships
(Hanihara & Ishida, 2005; Harris & Rathbun, 1991). This method reveals how tooth size is
apportioned across the dentition, thereby detecting differences and similarities between
populations.
Applications of quantitative and population genetic models have helped to develop import-
ant insights into population structure. In particular, R-matrix population models developed
for allele frequencies and applied to phenetic (or phenotypic) data (Konigsberg, 2006; Releth-
ford, 2012) are viewed by some as better than traditional model-free approaches (Hanihara,
2013; Stojanowski, 2005a, 2009). The genetic variation statistic (called FST) in R (relationship)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
360 Biological distance and historical dimensions of skeletal variation

10
–3.
10
–2.
GEG

10
–1.
0
Y
–.1

HES
SAQ
.90

QUR KHA

THE
0
1.9

GIZ
HAW
BAD
TAR
LIS
0
2.9

ABY HRK
2.00 NAQ

1.00

GRM
N

.00

–1.00

–2.00

1.70

.70

–.30

X –1.3
0

–2.3
0

Figure 9.2 Three-dimensional ordination of dental divergence (MMD) values from Neolithic
(GRM, Gebel Ramlah), Predynastic (BAD, Badarian; NAQ, Naqada; HRK, Hierakonpolis),
Early Dynastic (ABY, Abydos), Middle Kingdom (THE, Thebes), New Kingdom (QUR,
Qurneh), and Roman (HES, El Hesa; KHA, Kharga) Upper Egypt; and Early Dynastic
(TAR, Tarkhan), Old Kingdom (SAQ, Saqqara), Middle Kingdom (LIS, Lisht), Late Dynastic
(GIZ, Giza), Ptolemaic (GEG, Greek Egyptian), and Roman (HAW, Hawara) Lower Egypt.
(From Irish, 2006; reproduced with permission of author and John Wiley & Sons, Inc.)

matrices provides an important analytical tool to analyze genetic variation, especially with
regard to genetic distances relating to patterns of relationships between skeletal series within
a regional mating network. The statistic ranges from 0, where the samples are identical
genetically and engaged in considerable migration and genetic exchange, to 1, where all
samples are genetically distinct and have no genetic exchange. That is, if two populations are
exchanging mates frequently, they will express smaller phenotypic or genotypic differences
than two populations exchanging mates infrequently. Over time, changes in these values
represent changes in degree of mate exchange, patterns of migration, and gene flow in
general (Table 9.1).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.1 Introduction 361

Table 9.1 Temporal change in genetic diversity (FST) in Spanish Florida (From Stojanowski,
2009)
Period FST value Interpretation
Precontact 0.018 Significantly different from 0, indicating genetic exchange but with
(pre-AD 1550) limited gene flow
Early Mission 0.024 Increased values reflect somewhat increased genetic isolation, limited
(AD 1600–1650) gene flow, and likely little change in mating patterns in comparison
with the precontact populations
Late Mission 0.002 Dramatic decline in genetic variance indicates decrease in biological
(AD 1650–1704) structure and no isolation by distance. The record suggests increased
long-distance migration and gene flow during a time of population
decline and collapse

Biodistance analysis is problematic largely owing to the multifactorial nature of the traits
being studied. As with phenotypic variation in general, skeletal and dental traits are influ-
enced by a combination of intrinsic (genetic) and extrinsic (environmental) factors. Nowhere
is this more apparent than in skeletal tissue, which undergoes considerable modeling and
remodeling over the course of the lifetime of a person. Dental enamel – the hard tissue
containing most variation studied by biological anthropologists – does not remodel, but
nonetheless is complex and influenced during formation by a variety of factors (see Chapter 2).
It is not surprising, therefore, that skeletal and dental heritability – one estimate of the influence
of genetics – is so imprecisely understood (for some estimates of heritability, see Carson, 2006;
Goose, 1971; Hughes et al., 2000; Hughes & Townsend, 2013; Potter et al., 1976; Rizk et al.,
2008; Sjøvold, 1984, 1995; Townsend, 1992; Townsend et al., 2003). Thus, the application of a
biological population model – which assumes at least some degree of genetic control of traits
and that the distances based on these traits are directly proportional to those derived from
gene frequencies – is difficult to verify in human populations drawn from archaeological
settings (Konigsberg, 2006; Saunders, 1989; Scott, 2008; Stojanowski & Schillaci, 2006). On
the other hand, genetic data sets from twin and family studies reveal a highly heritable
component as it pertains to dental morphology (Hughes & Townsend, 2013; Scott, 2008).
Heritability remains an important focus of study, but developmental aspects of variation play
an increasingly important role in understanding the underlying biology of dental morphology
(Hunter et al., 2010).
Unlike DNA, neither metric nor nonmetric traits bear a one-to-one correspondence with the
genome of an individual. For example, the humeral septal aperture, a frequently used post-
cranial nonmetric variant for biodistance analysis, has a high negative association with
robusticity. Similarly, spondylolysis has often been assumed to be a population genetic
marker. However, spondylolysis is strongly influenced by activity load on the lower back
(see Chapter 5). Consistency of findings with other lines of evidence (e.g., archaeological or
historical), especially when care is taken to screen out traits influenced by activity or
representing biomechanical adaptations (Heathcote, 1994), suggests that nonmetric traits
can provide important insight into biological relatedness. Minimally, these traits are import-
ant for testing nonrandomness in skeletal series (Saunders, 1989; Spence, 1994). Familial

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
362 Biological distance and historical dimensions of skeletal variation

studies of modern humans (Saunders & Popovich, 1978; Sjøvold, 1995), laboratory mice
(Grüneberg, 1952), and rhesus monkeys (Cheverud & Buikstra, 1981a, 1981b, 1982) support
this assessment. Phenotypic anthropometric analysis seems to provide the same results as for
quantitative genetic analysis (Konigsberg & Ousley, 1995). In assessing relationships, it does
not matter much that heritability of traits may be low. Rather, what matters is that environ-
mental variation is random with respect to the traits being analyzed.
The role of biodistance analysis in addressing archaeological questions and in providing a
context for other bioarchaeological research is central to most of the topics considered in this
book. This chapter discusses the important contributions of biodistance analysis to biohisto-
rical questions, with regard to both diachronic and synchronic perspectives on variation.
Folded into this discussion are the ways in which these data are used to provide contextual
information for interpreting other skeletal and dental variation, some of which are highlighted
in this book (e.g., infectious disease). Although virtually every area of the skeleton and
dentition has been considered to one extent or another, the preponderance of findings are
based on cranial and dental remains (for a discussion of use of postcranial remains, see
Donlon, 2000; Finnegan, 1978; Saunders, 1978).
As reflected in the explosion of investigations of DNA, both ancient and modern, since the
mid-1990s, paleogenetics has become central for understanding biological relatedness and
population history in a wide range of settings (Brandt et al., 2013; Brown & Brown, 2013;
O’Rourke, 2010; Raff et al., 2011; Relethford, 2010; Stone, 2008). Owing to the rapid onset and
degradation of DNA following death, most techniques used for documenting DNA from living
subjects are not appropriate for study of ancient remains, largely because of the very low
concentration and presence of mostly short, damaged fragments of DNA (Brown & Brown,
2011; Kaestle, 2010; O’Rourke, 2000; Pääbo et al., 2004; Stone, 2000; Willerslev & Cooper,
2005). However, polymerase chain reaction (PCR) and the more recent next generation
sequencing (NGS) technologies have opened a new window onto documenting and interpret-
ing genetic variation in archaeological remains. In particular, this technology has made it
possible to amplify single molecules, rendering them available for detailed analysis. Although
interesting and revolutionary when first reported (Doran et al., 1986; Pääbo, 1985), because of
what we now know about contamination and preservation in various environments, much of
the record of early discovery of aDNA has either been disproven or remains unsubstantiated.
However, important advances in improvement of recovery and especially accounting for
contamination are beginning to reveal the full strength of the study of aDNA for documenting
intra- and inter-population relatedness in past populations (Adler et al., 2011; Brandt et al.,
2013; Raff et al., 2011; Rasmussen et al., 2014). On the other hand, this is not to say that the
issues of contamination are solved. Rather, it remains a matter of considerable importance as
researchers determine the authenticity of aDNA (Brown & Brown, 2013).

9.2 Classes of biodistance data


There are three classes of biodistance data considered in bioarchaeological research:
skeletal and dental metrics, skeletal and dental nonmetrics, and aDNA. Metric traits are
continuous variables obtained from linear measurements or indices derived from measure-
ments (e.g., lengths, breadths, subtenses, radii), which are used to characterize size and shape
of skeletal elements, especially the cranium (Table 9.2). Advances in morphometric analysis

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.2 Classes of biodistance data 363

Table 9.2 Cranial measurements used in biodistance analysis (See definitions by


Buikstra & Ubelaker, 1994)
Measurement Measurement
Maximum cranial length Interorbital breadth
Maximum cranial breadth Frontal chord
Bizygomatic diameter Parietal chord
Basion–bregma height Occipital chord
Cranial base length Foramen magnum length
Basion–prosthion length Foramen magnum breadth
Maxilloalveolar breadth Mastoid length
Maxilloalveolar length Chin height
Biauricular breadth Mandible body height
Upper facial height Mandible body breadth
Minimum frontal breadth Bigonial width
Upper facial breadth Bicondylar breadth
Nasal height Minimum ramus breadth
Nasal breadth Maximum ramus breadth
Orbital breadth Mandible length
Orbital height Mandible angle
Biorbital breadth

involving digitizing nodes on the cranium and other skeletal elements to produce three-
dimensional images provides increased precision of quantification (McKeown & Jantz, 2005;
Richtsmeier et al., 1992; Ross & Ubelaker, 2010). Nonmetric traits are discrete or quasi-
continuous anatomical entities often expressed as gradations from absence to full expression
(e.g., metopism, incisor shoveling).
A high degree of variation exists in cranial shape in human populations, which is heavily
influenced by the manner in which the teeth and jaws are used in masticatory and nonmasti-
catory functions (see Chapter 7). Nevertheless, heritability studies reflect the considerable
genetic and epigenetic contributions to cranial shape (Carson, 2006; Cheverud et al., 1979;
Sjøvold, 1984; Stefan & Chapman, 2003). As such, craniometric data have been a key resource
in identifying cultural-historical associations and in characterizing relatedness in a wide range
of settings (Droessler, 1981; Hanihara et al., 2008; Heathcote, 1986; Hemphill & Mallory, 2004;
Howells, 1973, 1989; Hubbe et al., 2010; Jantz, 1973; Key, 1983; Matsumura, 2006, 2011;
Pietrusewsky, 1994, 2006, 2008, 2010; Powell & Neves, 1999; Ross & Ubelaker, 2010; Schillaci &
Stojanowski, 2005; Sciulli, 1990; Spence, 1974a, 1974b, 1994; Stefan, 1999; Varela et al., 2008;
and many others). More than 200 cranial nonmetric traits have been identified in humans
(Hauser and De Stefano, 1989) and include four primary types – ossicles (small bones) within
cranial sutures (e.g., lambdoidal ossicles); hyperostotic traits or unusual skeletal proliferations
such as bridges of bone (e.g., atlas bridging); hypostotic traits involving ossification deficiencies
(e.g., metopic suture); and foramen variation (e.g., single or double supraorbital foramen)
(Buikstra & Ubelaker, 1994; Ossenberg, 1970; Saunders, 1989; Tyrrell, 2000) (Figure 9.3).
Similarly, analyses of tooth size, shape, and morphology are also important for understand-
ing population structure and history, and because of the far greater degree of genetic

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
364 Biological distance and historical dimensions of skeletal variation

canalization of tooth shape and size, teeth may provide stronger forms of data for many types
of biodistance research. Built on earlier work by Albert Dahlberg (1956, unpublished), there
are some 30 standard nonmetric dental traits presently used in biodistance analysis (Scott,
2008; Scott & Turner, 1997; Turner et al., 1991) (Table 9.3). These traits include morphological
variants on tooth crowns (e.g., incisor shoveling, Carabelli’s trait) as well as root variation
(e.g., lower molar root number) (Figure 9.4). There is a strong genetic signal to these traits
implied by studies of living populations (Biggerstaff, 1970; Harris & Bailit, 1980; Hughes &
Townsend, 2013; Lundström, 1963). Thus, at its most simple level, populations containing a
high frequency of rare traits have a high degree of biological relatedness (Alt et al., 2013).
Coupled with the conservative nature of their evolution, dental traits and other dental
variation have proven important for biodistance analysis (Brace & Hinton, 1981; Calcagno,
1989; Dahlberg, 1951, 1963; Greene, 1982; Griffin et al., 2001; Hanihara et al., 2008; Harris &
Rathbun, 1991; Hemphill, 1991; Irish, 2005; Lukacs, 1983, 1989; Lukacs & Hemphill, 1991,
1993; Matsumura, 2006; Pilloud & Larsen, 2011; Sciulli, 1998; Sciulli et al., 1984; Scott &
Turner, 2006; Stojanowski, 2005b; Turner, 1991, 1993b; Ullinger et al., 2005; and many
others).
The aforementioned PCR and later technology have made it possible to extract, amplify,
and identify DNA from both the nuclear and mitochondrial genomes recovered from ancient
tissues. Controlling for a variety of issues, especially contamination by modern DNA, it has
become possible to identify the unique patterns of DNA in ancient remains associated with

(a)
1. Metopic suture complete

2a. Supraorbital notch present,


< 1/2 occluded by spicules

2b. Supraorbital foramen present

3. Infraorbital suture complete


5. Zygomaticofacial foramen
multiple, small
4. Multiple infraorbital foramina
two, distinct

Figure 9.3 Cranial nonmetric traits. (From Buikstra & Ubelaker, 1994; reproduced with
permission of the authors and the Arkansas Archaeological Survey.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.2 Classes of biodistance data 365

(b) 6. Parietal foramen


present on parietal bone

7c. Bregmatic bone present

7e. Apical bone present

7b. Coronal ossicle present


7f. Lambdoid ossicles present

7d. Sagittal ossicles present

7a. Epipteric bone present


7b. Coronal ossicle present

7f. Lambdoid ossicles present

7i. Parietal notch bone present

7g. Asterionic bone present

7h. Ossicle in occipitomastoid suture

Figure 9.3 (cont.)

both genomes (Willerslev & Cooper, 2005). Nuclear DNA provides the key source of the unique
genetic profile of an individual. This pattern presents itself in a series of tandemly repeated
nucleotides known as short-tandem repeats (STRs), which are a series of two nucleotides
repeated many times (Brown & Brown, 2011). In archaeological contexts, DNA typing based
on STRs has been successful in identification of family groups (Haak et al., 2008; Keyser-
Tracqui et al., 2003; Ricaut et al., 2006). Comparison with biodistance analysis based on
cranial and dental discrete trait analysis in an archaeological series from northern Mongolia
revealed concordance between the genetic evidence based on STRs and discrete traits (Ricaut
et al., 2010). Unfortunately, because there is only one diploid copy of nuclear DNA in each cell
and a high degree of DNA degradation, it is extremely unlikely that the STRs will be preserved
in most circumstances (Brown & Brown, 2011). In contrast, mitochondrial DNA (mtDNA) has
a very high copy number – about 8000 identical copies of the mtDNA genome are present in a
cell – which greatly increases the chance of finding diagnostic DNA, even if degraded. Further,
mtDNA is inherited through the mother only, facilitating the documentation of maternal

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
366 Biological distance and historical dimensions of skeletal variation

Table 9.3 Dental traits used in biodistance analysis (See definitions by Turner et al.,
1991; Scott & Turner, 1997)
Dental trait Teeth
Winging Upper first incisor
Shoveling Upper incisors, canine, lower incisors
Labial convexity Upper incisors
Double-shoveling Upper incisors, canine, first premolar, lower incisors
Interruption groove Upper incisors
Tuberculum dentale Upper incisors, canine
Canine mesial ridge Upper canine
Canine distal accessory ridge Upper and lower incisors
Premolar mesial and distal Upper premolars
accessory cusps
Tricuspid premolars Upper premolars
Distosagittal ridge Upper first premolar
Metacone Upper molars
Hypocone Upper molars
Cusp 5 (metaconule) Upper molars
Carabelli’s trait Upper molars
Parastyle Upper molars
Enamel extensions Upper premolars and molars
Premolar root number Upper premolars
Upper molar root number Upper molars
Radical number All teeth
Peg-shaped incisor Upper second incisor
Peg-shaped molar Upper third molar
Odontone Upper and lower premolars
Congenital absence Upper premolars and molars, upper and lower second
premolars, upper and lower third molars
Premolar lingual cusp Lower premolars
variation
Anterior fovea Lower first molar
Groove pattern Lower molars
Cusp number Lower molars
Deflecting wrinkle Lower first molar
Distal trigonid crest Lower molars
Protostylid Lower molars
Cusp 5 Lower molars
Cusp 6 Lower molars
Cusp 7 Lower molars
Canine root number Lower canine
Tome’s root Lower first premolar
Lower molar root number Lower molars
Torsomolar angle Lower third molar

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.2 Classes of biodistance data 367

(a)

(b)

(c)

Figure 9.4 Three examples of dental nonmetric traits: grades of labial convexity (a),
mandibular first molar anterior fovea size (b), and maxillary second incisor shoveling (c).
(From Turner et al., 1991; photographs by Scott Haddow; reproduced with permission of
authors and Wiley-Blackwell, a division of John Wiley & Sons, Inc.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
368 Biological distance and historical dimensions of skeletal variation

lineages and showing maternal relationships. Specific patterns of mtDNA are identified on the
basis of haplogroups – groups of mitochondria sharing a specific set of markers (Raff et al.,
2011). Thus, individuals or populations having the same haplogroup are more closely related
biologically than individuals or populations having different haplogroups.
Some mtDNA markers are unique to specific geographical regions as well as to single
populations (Brown & Brown, 2011; Kaestle, 2010; O’Rourke & Raff, 2010). In this regard,
analysis of mtDNA from archaeological contexts has provided important insights into
large regional patterns of population history, such as in the peopling of the Americas
(O’Rourke & Raff, 2010; Raff et al., 2011; Reich et al., 2012), the beginning of human
population expansion across the Pacific (Knapp et al., 2012), the impact of Aztec imperi-
alism on local populations in Mesoamerica (Mata-Miguez et al., 2012), population history
in Neolithic and post-Neolithic Europe (Brandt et al., 2013), and in a wide variety of other
geographical and cultural contexts (Adachi et al., 2009; Alzualde et al., 2007; Bolnick &
Smith, 2007; Carnese et al., 2010; Deguilloux et al., 2011; Fan Zhang et al., 2010; Forgey,
2011; Kaestle, 1995; Kaestle et al., 1999; Kemp et al., 2009; Lertrit et al., 2008; Malhi et al.,
2003; Moraga et al., 2005; Raff et al., 2011; Ricaut et al., 2006, 2010; Shimada et al., 2004;
Stone & Stoneking, 1993).

9.3 Biohistorical issues: temporal perspectives


9.3.1 Temporal continuities and discontinuities: identifying biological lineages
and histories
Assessment of cultural transitions and temporal changes in biological structures and the cause
of these changes are key issues in the study of past populations. For example, the Late Bronze
Age–Early Iron Age transition in the southern Levant is among the most studied in the region,
including changes in material culture, architecture, and mortuary practices. Traditionally, the
argument has been that these changes were due to invasion by foreign intruders (Redford,
1992). However, some argue that indigenous sociopolitical transformations were the primary
force in explaining changes (Dever, 1997). In order to test the hypothesis that cultural and
social change was an internal event, Ullinger and coworkers (2005) undertook a biodistance
analysis of a series of dental remains from Tell Dothan and Tell ed-Duweir (Lachish) repre-
senting the Late Bronze Age–Early Iron Age transition. Although dental trait complexity was
different between the two series, the analysis revealed no clear distinctions between them,
thus indicating broad biological continuity. Therefore, it is unlikely that the cultural and
social changes that characterize this setting originated with invasion and replacement by
some external group or groups.
Of particular importance is a consideration of how much biological change (e.g., tooth size
reduction or cranial shape change) is due to migration and gene flow and how much is due to
intrinsic developments, such as local adaptive transitions. If biological change is due primarily
to introduced sources of variation from elsewhere, then can the source of this external
influence be identified? This question has been addressed by researchers in a widespread
number of settings, especially in regard to identifying temporal and geographical continuities
and discontinuities and their meaning in population history and in social and cultural
transformations.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.3 Biohistorical issues: temporal perspectives 369

Nubia
The traditional model of population structure and history in the Nile River valley is based on
the assumption that a series of migrations of “racially” distinct populations replaced one
another at various times. The model posits that in the past there were two racial types living in
the Nile River valley. These include a “Caucasoid” type in the Upper Valley and a “Negroid”
type in the Lower Valley. Strouhal (1971) views biological changes in the region as resulting
from the arrival and greater influence of one type over another at various times (see discus-
sion by Armelagos et al., 1981). Thus, the traditional model invokes population movement
and gene flow to account for biological change in this region. From this perspective, the
reduction in craniofacial robusticity found in X-group populations is interpreted as signaling
the arrival of a “Negroid” population (Batrawi, 1946; and see Chapter 7).
In order to evaluate the traditional model of population replacement in Nile Valley history,
Greene and colleagues (Greene, 1967, 1972, 1982; Greene & Armelagos, 1972; Greene et al.,
1967) analyzed a battery of dental nonmetric traits within samples from Wadi Halfa, Sudan.
Univariate statistical comparisons of nonmetric dental traits revealed no significant differ-
ences between Meroitic and X-group samples. Based on these results, they concluded that
population replacement was unlikely in this setting, “but instead represent(s) a continuum of
(temporally) successive local Mendelian breeding populations” (Greene, 1982:76). Although
suggestive, univariate statistical analyses are problematic, especially because such analyses
incorrectly assume that individual traits are independent. In order to provide a more convin-
cing argument for demonstrating biological continuity, Greene (1982) employed the MMD
distance statistic to dental nonmetric traits to test the hypothesis that inter-population
biological differences are not significant. This analysis revealed extremely low divergence
values, and thus indicates no significant change in biological groupings, or the virtual lack of
infusion of peoples with different biological histories into the region. The implications of this
study are important because they lend strong support to the argument that biological change
in Nubia was due to local circumstances and not gene flow.
Turner and coworkers (Irish & Turner, 1990; Turner & Markowitz, 1990) reassessed Greene
and coworkers’ continuity model of Nubian population history. Although lauding the studies
of Greene and others for using a non-racial, non-typological approach to Nubian population
history, they argued that extra-regional sources of variation have been insufficiently con-
sidered, especially sources that may explain temporal shifts in craniofacial morphology in this
region. Implicit in the work by Greene and others is the assumption that population continuity
in Nubia extends at least as far back as the late Pleistocene (c. 12 000 yBP). Turner and
coworkers contended that continuity can only be claimed if non-Nubian populations are also
considered in statistical analyses of dental trait variation in this region.
Turner and coworkers included in their analysis additional Nubian dentitions from the late
Pleistocene Upper Stone Age, Meroitic, X-group, Christian period, and historic-era European
samples. Computed MMD values, modified for small samples (Green & Suchey, 1976) and
tested for significance (Sjøvold, 1973), reveal few significant differences in comparing the
Meroitic, X-group, and Christian periods, thus confirming Greene’s earlier conclusions
regarding population continuity. However, significant differences between the Pleistocene
and later groups were clearly identified (e.g., frequency of incisor shoveling and three-rooted
mandibular molars). Based on the apparent temporal discontinuity between the Pleistocene
and later populations, Turner and Markowitz (1990) hypothesized that the ancestry of recent

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
370 Biological distance and historical dimensions of skeletal variation

Nubians did not derive from local late Pleistocene populations, and that a population
replacement event occurred during the Holocene in Nubia. The origin of these later popula-
tions is unclear, but based solely on dental traits, they argued that populations north of Nubia
containing European and Near Eastern traits are the most likely sources.
In an effort to identify other possible sources of variation, Irish and Turner (1990) compared
their sample of Nubian dentitions (late Pleistocene to Christian period) with historic-period
dentitions from a West African group – the Ashanti. Univariate and multivariate (MMD)
statistical treatment of these samples revealed strong similarities between modern West Africans
and late Pleistocene Nubians. As with previous studies, later Holocene dentitions were found
to be very similar. The late Pleistocene and modern West Africans are strongly divergent from
the Meroitic, X-group, and Christian-period Nubians. Thus, they argued that there is a
population discontinuity between the late Pleistocene and Holocene populations in Nubia,
with the former sharing biological affinities with West Africans. Irish and Turner (1990)
suggested that the discontinuity can be explained by high rates of violence and decline (or
possibly extinction) in late Pleistocene (Mesolithic) Nubian foragers, which may have left
them susceptible to invasion or “genetic swapping” by other groups from West Africa
(Turner & Markowitz, 1990).
In a more detailed analysis containing remains from Nubia not included in previous studies,
Irish (2005) performed additional multivariate analysis to document patterns of discontinuity
and continuity inferred from biodistance analysis. His comprehensive analysis revealed a
clear discontinuity between the late Paleolithic Jebel Sahaba series (14 000–12 000 yBP) and
all later series. That is, the late Paleolithic series displays higher frequency of some specific
dental traits than all other (later) series, such as in UI1 shoveling, six-cusped LM1, and LP1
Tomes’s root. Irish concluded that the Jebel Sahaba population is likely not ancestral to any
later regional population. However, the entire record of Final Neolithic through Christian
periods shows homogeneity in dental variation, suggesting regional population continuity
from at least c. 5600 yBP onward (Irish, 2012). It is during this later time frame that much of
the craniofacial reduction described by Carlson & Van Gerven (1977, 1979; Van Gerven et al.,
1995) took place in Nubia. Thus, although the high degree of robusticity displayed in pre-
Neolithic populations is apparently not the ancestral morphology for the region, there still
remains the presence of significant craniofacial size reduction in the post-Mesolithic popula-
tions, changes taking place within a single population evolving over time in situ (see
Chapter 7). Thus, viewed in the context of changes in diet, food preparation, and technology
in general, these craniofacial modifications were likely due to reduced mechanical loading of
the jaws and teeth in later times. That said, the Nubian populations did not live in isolation,
and the record presented through biodistance analysis by Irish is consistent with mtDNA
studies of modern people showing evidence for gene flow especially from Egypt (Buzon, 2006;
Irish & Konigsberg, 2007) as well as the dynamic nature of population movement in the Nile
Valley corridor in general (Irish, 2000, 2006; Zakrzewski, 2007) and during specific times and
places (Buzon, 2011).

Southern Africa
Early models accounting for the origin of population variation among native peoples of
southern Africa argued for large-scale immigration and complex processes associated with
isolation and development of distinctive populations (Ginter, 2011; Hausman, 1984;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.3 Biohistorical issues: temporal perspectives 371

Morris, 1992; Rightmire, 1970, 1976). A range of anthropological and historical evidence
suggests that cultural and biological characteristics of modern native peoples may have arisen
from in situ processes associated with the transition from foraging to nomadic pastoralism
at about 2000 yBP (Ginter, 2011). Morphometric analysis of archaeological crania addresses
the issue of evolution and population history in this very dynamic region (Ginter, 2011;
Hausman, 1984; Irish, 2013; Morris, 1992; and see Rightmire, 1970, 1976).
When first encountered by Europeans, native peoples of southern Africa were described
as consisting of two distinctive groups, including the San, who were relatively small in
body size and who followed an exclusively foraging lifeway. The Khoi, who were described
as taller and more robust than the San, pursued a pastoralist lifeway based on sheep and
cattle husbandry. Khoi pastoralists were widespread, but lived mostly in coastal areas (e.g.,
Eastern Cape Province) and in some major river valleys in the interior of southern Africa.
The origin of pastoralism in recent Khoi is unclear, and revolves around the question: Was
it due to a migration of people into the region or a diffusion of ideas? The consensus
drawn from ethnographic, linguistic, and archaeological data is that people migrated to
southern Africa c. 2000 yBP, bringing with them pastoralism (Deacon & Deacon, 1999; but
see Sadr, 2003).
Multivariate discriminant function analysis of craniometric traits confirms the biological
distinctiveness of Khoi and San peoples (Hausman, 1984; and compare with Morris, 1992;
Rightmire, 1970, 1976). Although separated from other South Africans, crania of these two
groups are morphologically similar, suggesting that they share a common origin. Based on her
analysis of cranial morphology, Hausman proposes that one segment of the population
adopted pastoralism, which contributed to “the biological separation of the groups”
(1984:268). Differences between Khoi and San marriage and kinship systems may have
contributed to this biological variation (Hausman, 1984). Exogamy among the highly mobile
San facilitates gene flow, and thus may explain the high degree of homogeneity across local
San groups within broad regions. Khoi are exogamous primarily along patrilineal clan lines
within individual tribes. These marriage patterns, combined with other practices (e.g., cross-
cousin marriage) would increase interbreeding and homogeneity within small populations.
These social behaviors would discourage interbreeding between Khoi and San, resulting in
greater genetic isolation and morphological differentiation between the two groups. Overall,
social behavior has enormous implications for patterns of gene flow and the way that these
patterns influence biological variation in South Africa (Morris, 1992).
Biodistance analysis of dental variation in sub-Saharan Africa reveals an overall uniform-
ity of variation (Irish, 1997, 2013), so much so that Irish (2013) identified an “Afridont” dental
pattern, including a series of high-frequency traits (e.g., Bushman canine, two-rooted
maxillary first premolar, two-rooted mandibular second molar). Moreover, the craniometric
findings suggest very little evidence for change before and after the shift from foraging to
pastoralism among prehistoric Khoi in the coastal region (Ginter, 2011). In contrast to
traditional models for origins of Khoi coastal pastoralists, this finding argues that population
replacement did not occur in this region. Rather, pastoralism was introduced to and adopted
by local populations (Hausman, 1984). Similarly, analysis of cranial, dental, and postcranial
measurements and cranial discrete traits using multivariate and univariate statistical analysis
of a large data set from the Eastern Cape Province reveals strong evidence for biological
continuity spanning the pre-herding and herding traditions in the region and a general

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
372 Biological distance and historical dimensions of skeletal variation

pattern of biological homogeneity (Ginter, 2011). Collectively, these analyses strongly favor
an in situ development of pastoralism in the southern African region.

European Neolithization and the spread of farming: demic or cultural diffusion, or both?
As with much of the rest of the world, the period of the late Pleistocene and early Holocene in
Europe was enormously dynamic, involving movement of people, their technology, and their
ways of acquiring food and other resources. Populations saw a transition from a lifeway
and subsistence economy based on nondomesticated to domesticated plants and animals,
representing a profound alteration in nutrition and lifestyle. Humans first began domesti-
cating animals (goats, sheep, pigs, and cattle) and plants (wheats and barleys) around
13 000–10 000 yBP in the Near East (Pinhasi et al., 2005; Smith, 1998; Zeder, 2008). By
8500 yBP, this Neolithic package had spread to Greece, and by 5000–6000 yBP to Scandinavia
and Great Britain (Burger & Thomas, 2011). Based on modern genetic evidence, a number of
authorities argue for a demic diffusion process whereby the spread of domestication was
occasioned by the migration and dispersion of early farmers from the Near East via Anatolia,
replacing indigenous hunter-gatherers throughout Europe (Barbujani et al., 1998; Cavalli-
Sforza et al., 1996). In contrast, the competing cultural diffusion model argues that while there
may have been population dispersion in some areas, in part originating from the Near East, it
was the movement of the idea of agriculture and domestication that explains the Neolithic
transition (Whittle, 1996; Zvelebil, 1986). The former should be represented by clines of
genetic variation across Europe, whereas the latter would predict similar kinds and frequen-
cies of genetic variants between the earlier foragers and later farmers of Europe (Deguilloux
et al., 2012; Pinhasi & von Cramon-Taubadel, 2009).
A wide variety of non-molecular and molecular data presents a contradictory picture
indicating evidence of both migration and replacement and indigenous adoption of farming
and domestication generally (Deguilloux et al., 2011; Fox, 1996; Harding et al., 1989;
Linderholm, 2011). Analysis of aDNA from a series of Neolithic individuals from Germany,
Austria, and Hungary revealed a substantial presence of sequence-specific genetic markers
particular to haplogroup N1a (25%), found only rarely in modern Europeans (0.2%) (Deguil-
loux et al., 2011; and see Pinhasi et al., 2012). This finding suggests that modern Europeans
may not be the descendants of the first farmers in the continent (Haak et al., 2005; and see
Pinhasi & von Cramon-Taubadel, 2009). Similarly, analysis of aDNA in Sweden provides a
genetic reconstruction showing that Mesolithic hunter-gatherers may have been replaced by
farmers moving into the region (Linderholm, 2011). Moreover, analysis of genetic diversity,
including markers in the Y-chromosome, appears to be best understood in relation to a single
population source in the Near East (Anatolia) during the Neolithic (Deguilloux et al., 2011).
In reality, the process of the adoption of domestication and especially farming was complex
and almost certainly involved different and region-specific circumstances across Europe. The
genetic discontinuities suggest that it is likely that all modern Europeans are not the direct
descendants of either Paleolithic hunter-gatherers or migrating Neolithic farmers. Thus, the
question of demic or cultural diffusion may not be the best one to ask. Rather, it would appear
from the complexity of the record that a much more interesting problem is to explore the wide
variation, such as that observed in modern and ancient DNA. This conclusion about popula-
tion history in Europe is indicated by the likelihood of periods of continuity and discontinuity
based on mtDNA (Brandt et al., 2013). Until complete diachronic data sets of autosomal aDNA

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.3 Biohistorical issues: temporal perspectives 373

become available, especially those that present evidence for continuity (or discontinuity), the
answers to questions about population origins and evolution – with or without domestication
in pre-Neolithic and Neolithic contexts – remain unclear.
Skeletal biodistance analyses undertaken by Schwidetzky and Rösing for the Neolithic/
Bronze Ages (Schwidetzky, 1967), Iron Age (Schwidetzky, 1972), Roman period
(Schwidetzky & Rösing, 1975), and Medieval and recent periods (Rösing & Schwidetzky,
1977, 1979, 1984; Schwidetzky & Rösing, 1982) also bear out the complex nature of European
population history. Their cluster analysis indicates two primary morphometric complexes –
Western and Eastern – with size gradients between the two. In the Western complex, cranial
length is greater than that in the Eastern complex; in the Eastern complex, cranial width and
bizygomatic breadth, are greater (Rösing & Schwidetzky, 1979). During the Early Middle Ages,
distinctions between the groups become less straightforward. Schwidetzky and Rösing sug-
gested that such a reduction in distinctiveness is due to increased population movement and
increased gene flow. In the following Late Medieval period, the distinctions between Eastern
and Western complexes become even less clear (Rösing & Schwidetzky, 1984).
Other cluster analyses of these European samples reveal little evidence of sharp differenti-
ation between groups during the Medieval period (Sokal et al., 1987), which contrasts with a
record of abrupt gene frequency boundaries in modern Europe (Barbujani & Sokal, 1990).
Sokal and coworkers (1987) contended that the apparent lack of clear definition between
groups in later European history may be related to many factors, including variability in
measurement accuracy among observers and definitions of characters, small sample size for
some sites, the nature of variables used in the study, and incorrect sex identification. Cranial
lengths, breadths, and facial size are developmentally plastic (see Chapter 7). Thus, some of
the variability documented in European populations may also be influenced by shifts in
technology or dietary practices and the craniofacial responses to these shifts. Long-term
contacts (e.g., Gothonic-Slavic groups), admixture (e.g., Turkic-Mongolian groups), wide-
spread migration (e.g., Italian groups), and random genetic drift may have influenced the
patterns of cranial variation seen in these samples. In summary, no clear groupings are
evident when looking at this very broad region.

Central and South Asia


Among the pressing concerns surrounding biological and cultural history in South Asia is the
origin, rise, and decline of the Harappan Civilization of the Indus River valley of Pakistan. The
Harappan Civilization arose and attained a peak of urbanization and intercommunity coord-
ination in the lower Indus Valley during the third millennium BC, and is the first large-scale
complex society on the Indian subcontinent (Chakrabarti, 1980; Fairservis, 1971). Marshall
(1931) argued that the Harappan Civilization developed from earlier pre-Harappan cultures
within the Indus Valley, whereas others argued that it was the product of “stimulus diffusion”
or migration from the West, especially Mesopotamia (Gadd, 1932; Gordon, 1958; Wheeler,
1968; and see discussions in Hemphill et al., 1991; Lukacs, 1989). Similarities in ceramics,
metallurgy, and other material culture in the archaeological record indicate connections
between Indus Valley populations and the populations of Mesopotamia, the Iranian Plateau,
and southern Central Asia. Other archaeological evidence, however, suggests that such
connections, if they existed at all, were indirect and unidirectional (reviewed by Hemphill
et al., 1991).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
374 Biological distance and historical dimensions of skeletal variation

Hemphill and coworkers (1991, 1996) employed multivariate statistical analyses of cranio-
metric, cranial nonmetric, and dental nonmetric traits for Indus Valley samples in an
effort to resolve problems in understanding biocultural change and population origins in
this exceedingly complex region. Cluster and principal components analyses of Harappan-
phase dentitions and skulls from Cemetery R37 at the site of Harappa provide two principal
findings. First, the differences between Neolithic and the later Chalcolithic dentitions from
the Indus Valley suggest a lack of population continuity between about 6000 and 4500 BC.
Second, a general similarity in traits between sites in the region after 4500 BC indicates the
presence of biological continuity during the rise, apogee, and decline of the Harappan
culture. Iron Age populations (c. 200 BC) are markedly different in pattern of dental
nonmetric variation, suggesting a discontinuity during later prehistory after the demise of
Harappan culture. Analysis of cranial nonmetric traits indicates that the source of this
discontinuity may be gene flow into the region from the Iranian Plateau in the Near East.
In summary, the results of this research document the lack of biological continuity from the
Neolithic to the period of Christianity. Rather, two discontinuities existed, the first between
6000 and 4500 BC and the second between 800 and 200 BC. These findings suggest that
although the Harappan culture was an indigenous development within the Indus Valley, the
region nevertheless interacted on a widespread level at various times, especially with
populations to the west.
Unlike the Indus Valley, the biological and archaeological history of central Asia is poorly
documented, in part because the region has long been envisioned as culturally barren, serving
primarily as a buffer between the Harappan culture to the east and the Mesopotamian
societies to the west (Wheeler, 1968). Archaeological research beginning in the 1970s and
1980s revealed that the central Asian borderlands were inhabited by various complex soci-
eties, including the Oxus culture, a culture that was associated with oases in Bactria and
Morgiana, located in Uzbekistan, Turkmenistan, and Afghanistan (Hemphill, 1998, 1999a,
1999b; Hemphill et al., 1996, 1998). This region may have played an important role in the
diversification and spread of Indo-Iranian languages into the Iranian plateau and Indian
subcontinent (Hiebert and Lamberg-Karlovsky, 1992). Based on archaeological evidence, and
especially material culture, some argue that the Bactrian culture is an eastward expansion of
the central Asian oasis society, originating in western Turkmenistan and northwestern Iran.
The presence of Harappan artifacts in Bactrian sites indicates to some authorities that the
origins of Bactrian culture are tied to the Indus Valley to the east.
Univariate, cluster, and MMD statistical analyses based on dental and craniometric traits
from the northern Bactrian oasis sites in Uzbekistan (2300–1650 BC), localities in Turkmeni-
stan, Iran, the Indus Valley, and other South Asian localities, reveal strong separation between
central and south Asians (Hemphill et al., 1996). Unlike the South Asian groups, Bactrian
samples contain low frequencies of reduced buccal-distal cusp (metacone) in the first maxil-
lary molar. This and other distinctions indicate no evidence of a biological link between
central Asian and Harappan populations. Thus, there appears to be no support for a model
positing that the sedentary Bronze Age settlements in Bactria owe their origin to Harappan
settlers from the east; neither is there evidence to suggest that Bactrian populations moved
eastward into the Indus Valley following the decline of Harappan culture. Overall, the rise of
the Oxus culture appears to have occurred locally without significant external influence,
either from the east or from the west.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.3 Biohistorical issues: temporal perspectives 375

One of the most sensational archaeological discoveries in Central Asia involved the recov-
ery of a series of some 300 mummies – naturally desiccated bodies – that displayed “Europoid”
or “Caucasoid” facial features, found in the Tarim Basin of the western Chinese Xinjiang
Uyghur Autonomous Region in sites dating to the Bronze Age (c. 1800 BC). In order to
document and explain the origins of these individuals, Hemphill and Mallory (2004) under-
took a comprehensive craniometric biodistance analysis of Tarim Basin crania and numerous
other crania from a range of localities in Central Asia, South Asia, and China. The analysis
revealed that the population, while distinctive, shows no biological relationship to suggest
that it owes its origins to “Europoid” populations of the Russian steppes. Instead, its origins
appear to have been multiple, and the closest affinities may be found to the west among the
inhabitants of the North Bactrian oasis of Uzbekistan and the Iranian Plateau, as well as to the
south among populations of the Indus Valley.

East Asia and Japan


Some investigators see a highly generalized pattern of variation in East Asian biological traits,
especially in the dentition (Hanihara, 1969). Patterns of dental morphological variation, as
revealed by the study of hundreds of East Asian dentitions from a wide variety of settings,
indicate a more complex picture of population history (Turner, 1979, 1985, 1986a, 1986b,
1987). Turner identified a “Sundadont” group and a “Sinodont” group associated with
respective southern and northern populations in East Asia. Based on MMD multivariate
statistical treatment of dental traits from both archaeological and living populations, Turner’s
Sundadont cluster includes Southeast Asian and Jomon (Japan) populations and the northern
cluster includes all northern Asian populations. Sundadont dental morphology is character-
ized by having generally weaker grades of trait expression, such as incisor shoveling and
double-shoveling, and presence of mandibular molars with four cusps rather than five or
more. In contrast, Sinodont populations possess higher frequencies of upper first premolars
and lower second molars with single roots and lower first molars with three roots. Overall, the
Sundadont populations show morphological simplification, whereas Sinodont populations
exhibit more complex teeth. Turner regarded the Sundadonts as ancestral to the Sinodonts
with the divergence taking place between the two sometime prior to 15 000 yBP in north-
eastern Asia.
An important outcome of Turner’s ambitious investigation of dental trait variation in East
Asia is insight into the population history of East Asia, including specific regions – Japan,
Southeast Asia, the Pacific, and the Americas. Two primary hypotheses account for the origins
of living Japanese populations, the dominant Japanese population inhabiting most of the
country of Japan, and the Ainu of the northern island of Hokkaido (see discussions
by Hanihara, 1985; Hanihara & Ishida, 2009; Hanihara et al., 2008; Nakahashi, 1993;
Pietrusewsky, 2013). Based on temporal change in cranial measurements, some argue that
populations represent a biological continuum beginning with the Neolithic Jomon (the
original population from Asia) through modern Japanese (Suzuki, 1969, 1981). This suggests
that virtually all biological change, including a marked increase in face height (Hanihara,
1985; Nakahashi, 1993) and tooth size (Brace & Nagai, 1982) is due to cultural and dietary
shifts, such as the introduction of rice agriculture. Proponents of this view argue that although
immigration from mainland Asia may have influenced craniofacial morphology, the general
effects of such population movement on prehistoric and later Japanese were negligible.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
376 Biological distance and historical dimensions of skeletal variation

A second hypothesis posits that biological changes in the Jomon–Yayoi transition at about
300 BC were due to immigration and gene flow from the Asian mainland during the late
Jomon period (Hanihara, 1994; Hanihara et al., 2008; Nakahashi, 1993). This dual structure
hypothesis argues that biological and cultural influence from the mainland spread quickly
throughout Japan, eventually involving all populations except those inhabiting Hokkaido and
the Nansei Island chain (Aikens & Higuchi, 1982; for archaeological context, see Chard, 1974).
Because of such profound and widespread gene flow, proponents of this model maintain that
the ancestry of living Japanese is mostly linked with mainland Asians, whereas the modern
Ainu populations of northern Japan are descendants of the founding Jomon populations
(Hanihara & Ishida, 2009).
The study of dental traits and size is consistent with craniometric and cranial discrete trait
analyses completed by various workers (Dodo, 1986; Hanihara, 1979, 1985; Howells, 1966,
1986; Nakahashi, 1993; Ossenberg, 1986; Pietrusewsky, 2000; Yamaguchi, 1982). Multivari-
ate analyses indicate closer biological links between Jomon and living Ainu than between
Jomon and living Japanese. Collectively, these analyses support the alternative point of view
that the biocultural discontinuity during the late Jomon period resulted from migration and
gene flow. These analyses emphasize that this discontinuity is not a precipitous population
replacement. The gradient of biological distances in nonmetric cranial traits from Siberians–
Kinki–Kanto–Ainu–Jomon suggests the presence of gene flow and hybridization between
mainland immigrants and indigenous Japanese during the Yayoi and following periods
(Ossenberg, 1986).
The dual structure hypothesis for the peopling of the Japanese Islands is supported by a
growing body of data. However, while Turner’s Sundadonty versus Sinodonty hypothesis
helped to clarify population differences between the Jomon and Yayoi, the record is providing
increasing evidence that Jomon did not derive from affinities in Southeast Asia. Rather,
skeletal evidence for East Asia (Hanihara and Ishida, 2009; Pietrusewsky, 1999) suggests that
the migration responsible for Jomon people moved first from Northeast Asia into Hokkaido,
then through the southerly islands of Japan. These results are supported by aDNA analysis of
Jomon people from Hokkaido, suggesting that they have population affinities in southwestern
Siberia and not Southeast Asia (Adachi et al., 2009).
Population history in Southeast Asia has long been known to be complex and to involve
considerable migration and movement of people and ideas. One model argues for a later
immigration of populations from southern China to Southeast Asia, bringing with them
agriculture and other elements of complex society (Bellwood, 2001; Howells, 1973). This
model indicates hybridization between Neolithic northern East Asians and local resident
Australo-Melanesians. Turner’s (1987) analysis of the dental record and association of simpler
(Sundadont) dental morphology in the region revealed a different population history and
independent trajectory of biological history, the results of which were said to argue for local
evolution of Southeast Asian populations, independent of northern Asian groups, especially
from China. His model contends that groups with the Sundadont pattern spread north and to
Japan (see earlier), Siberia and northeastern Asia, the Americas, and the Pacific.
An alternative multivariate analysis of tooth size and morphology reveals a very different
pattern of biological history, consistent with the traditional immigration model involving
migration from south China and hybridization (Matsumura, 2006, 2011; Matsumura &
Hudson, 2005) versus Turner’s regional continuity model (Pietrusewsky, 1994, 1999, 2006).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.3 Biohistorical issues: temporal perspectives 377

As did earlier workers, Matsumura and others observed that the preceramic hunter-gatherers
of Southeast Asia have morphology similar to Australian Aborigines and Melanesians,
namely relatively long, narrow cranial vaults and robust faces and jaws. In contrast, later
ceramic-equipped, farming populations display rounder crania, reduced facial-masticatory
architecture, and smaller teeth, similar in some respects to Neolithic and post-Neolithic
populations from northern Asia. In light of these differences between earlier and later South-
east Asian populations, Matsumura and Hudson’s (2005) reconstruction of population history,
based on analysis of metric and nonmetric variation in a series of 4000 dentitions from Asia,
Australia, and Melanesia, revealed similarity between preceramic Southeast Asia and
Australo-Melanesians and biological dissimilarity between preceramic and later Southeast
Asians (Figure 9.5). In this “two-layer” model, the later dissimilarity is due to gene flow and
admixture derived from immigration of groups from North Asia or East Asia, or both
(Matsumura, 2011).
The question then emerges: Which of the two hypotheses is correct, local development and
dispersal, or population displacement originating from elsewhere in Asia? In consideration of
other biodistance analyses, the answer to this question is likely more complex than either
model indicates. Large-scale craniometric biodistance analysis shows two broad groupings of
prehistoric populations, including (1) Australia, Tasmania, New Guinea, and Melanesia, and
(2) East/North Asia, Southeast Asia, Polynesia, and other areas of Oceania (Pietrusewsky,
2005) (Figure 9.6). Given the distinctiveness of the clusters, they likely relate to two broad
origins of peopling of this vast region of the world. Within the second cluster, East/North Asia
and Southeast Asia are separated. This distinction may provide evidence for long-term in situ
evolution in each of these two regions and not replacement of ancient Southeastern Asian
populations via immigration from East/North Asia (Pietrusewsky, 2005, 2006).
However, genetic (PCR) analysis of mtDNA from teeth of 30 individuals from the Noen
U-Loke (400 BC–AD 500) and Ban Lum Khao (1200–400 BC) sites provides preliminary
evidence linking early agriculture Bronze–Iron Age populations from northeastern Thailand
with modern Thai and other East Asian populations (Lertrit et al., 2008). In the regional
context, inhabitants of these two settings present strong genetic similarity indicating they are
from the same gene pool of this region of Southeast Asia. Moreover, the close genetic affinity
with modern Thai indicates the strong possibility for a genetic continuity having time depth
extending back millennia. On the other hand, the presence of sequences shared with popula-
tions from southern China provides a complex picture of gene flow, and certainly does not
rule out demic diffusion from southern China into Southeast Asia (Lertrit et al., 2008). While
not a replacement process, this record suggests the important role of gene flow in the
development of modern populations in the region.

Coming to America: New World population origins and diversification


The origins and evolution of native populations in the Americas has been a point of discussion
and debate for well over a century, beginning with the founders of American archaeology and
physical anthropology (Larsen, 1985; Meltzer, 2009; Powell, 2005). Based on the strong
anatomical similarity between Native New World and Asian populations, especially in dental
characters (e.g., both share high frequencies of incisor shoveling), the origin of Native
Americans is unequivocally traced to Northeast Asia (Siberia) with immigration taking place
across the now-submerged Bering land bridge to present-day Alaska (Dahlberg, 1951, 1963;

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
378 Biological distance and historical dimensions of skeletal variation

Buriats
Urga Mongolians
East Asians
Northern Chinese
Chifeng Chinese
Koreans
Kofun Japanese
Tokyo Japanese
Kamakura Japanese
Edo Japanese
Neolithic Southern Chinese
Dayak
Laotians
Lesser Sunda & Java Islanders
Early Holocene Laotians
Ban Kao
Malay
Philippines
SE-Asians Sumatra Islanders
Chinese Thailanders
Thailanders
Yayoi
Anyang Chinese
Dong Son Vietnamese
Vietnamese
Prehistoric SE-Asians Negritos
Australians Melanesians
Guar Kepah
Jomon
Amami-Okinawa Islanders
Non Nok Tha
New Britain Islanders
Gua Cha
Loyalty Islanders
Andaman Islanders
Bac Son Vietnamese
Hokkaido Ainu
Da But Vietnamese
Australian Aborigines

Sakhali Ainu
Figure 9.5 Dendrogram of cluster analysis for prehistoric and historic Southeast Asian
populations, depicting relative genetic affinities between various populations. (From
Matsumura & Hudson, 2005; reproduced with permission of authors and John Wiley &
Sons, Inc.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.3 Biohistorical issues: temporal perspectives 379

Tonga-Samoa
Hawaii
Guam
Rapa Nui
Gambier Is.
Marquesas Is.
Society Is.
Tuamotu Arch.
Chatham Is.
Sumatra
Bomeo
L. Sundas
Java
Sulawesi
Sulu
Cambodia-Laos
Philippines
Vietnam
Bachuc
Thailand
Burma
Kanto
Tohoku
Kyushu
Taiwan
Hainan Is.
Korea
Ryukyu Is.
Atayal
Shanghai
Hangzhou
Nanjing
Chengdu
Hong Kong
Manchuria
Ainu
Mongolia
New Zealand
S. Moluccas
Fiji
Biak Is.
Solomon Is.
New Ireland
Caroline Is.
Marshall-Kiribati
Sepik R.
D’Entrecasteaux Is.
Dawson St.
Vanuatu
New Britain
Santa Cruz
Loyalty Is.
New Caledonia
Fly R.
Purari Delta
Admiralty Is.
Murray R.
New South Wales
Queensland
N. Territory
W. Australia
Swanport
Tasmania

Figure 9.6 Dendrogram of cluster analysis for prehistoric Asian and Pacific populations, based on craniometric
variables. (Pietrusewsky, 2005; reproduced with permission of author and RoutledgeCurzon.)

Harper & Laughlin, 1982; Hrdlička, 1912; Turner & Bird, 1981). A large body of archaeological
and genetic evidence indicates that the immigration of Old World peoples to the New World
took place by 15 000 years ago and likely earlier (Meltzer, 2009; Raff et al., 2011). Moreover,
although the routes and number of colonization events are subject to considerable debate,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
380 Biological distance and historical dimensions of skeletal variation

there is a consensus that the original founders migrating to North America derived from a
sampling of populations from Asia prior to the last glacial maximum (Raff et al., 2011).
With the development of molecular genetic methods for accessing and characterizing aDNA
in archaeological skeletal samples, this finding is now well confirmed and has facilitated more
direct ways of testing hypotheses regarding ancestral – descendant relationships for the
Americas generally (Adachi et al., 2009; Chatters et al., 2014; Kemp et al., 2007; O’Rourke
& Raff, 2010; Raff et al., 2010, 2011; Rasmussen et al., 2014; Schurr, 2004) and regionally. For
example, the peopling of the Aleutian archipelago was characterized as a population discon-
tinuity involving replacement of the original founding population (Paleo-Aleuts) by later
immigrants (Neo-Aleuts) by AD 1000 (Hrdlička, 1945). Based on study of aDNA and other
lines of evidence (Coltrain et al., 2006; Smith et al., 2009), the peopling of the region involved
at least two migrations with later groups coexisting with the earlier founding population. On
the other hand, the close genetic relationship between Paleoamerican pre-Holocene skeletons
and native North Americans indicates biological continuity between these early groups and
indigenous populations (Chatters et al., 2014; Rasmussen et al., 2014).
Analysis of dental traits by Turner (1985, 1986a, 1986b) indicates that Native Americans
are mostly derived from the Sinodont group of Northeast Asia. His extensive study – based on
some 9000 individuals – made the case that Native American populations can be further
subdivided into three groups, the Eskaleut (Eskimos and Aleut of the Arctic), Nadene
(Athapaskans, Tlingits, Haida, and Eyaks), and Macro-Indian (all other Native North, Central,
and South Americans) (Greenberg et al., 1986; Turner, 1986a, 1986b, 1987). Based on the
assumption that dental traits evolve at a constant rate, Turner used MMD distance statistics to
calculate a “dentochronology” for the three groups migrating separately from the Old World:
an initial Paleoamerican population divergence taking place 14 000 yBP between the
founding Sinodonts and the subsequent migration of the first migrant group (Macro-Indians);
a second split and migration along the southern coast of the Bering land bridge with
populations (Eskaleut) migrating at about the same time; and a slightly later split representing
a third population migration (Nadene). Turner’s tripartite model suggests that these migrations
were separate and were completed by 9000 yBP. Greenberg and coworkers (1986:485) argued
that there is “a remarkably good fit” between Turner’s tripartite dental groupings, linguistic
divisions of Native American languages, and genetic evidence of living populations for
inferred divergence and migration dates.
The concordance of the three lines of evidence – dental, linguistic, and genetic – has met with
strong criticism. For example, Turner (1985) assigned prehistoric dentitions in the Aleutian
region to the Aleut category, which assumes that the past and present population distributions
are the same. This assumption is unlikely, especially in consideration of the remarkable
time depth and morphological and genetic diversity of the populations in this setting
(Meltzer, 1993, 1995; Raff et al., 2010; Smith et al., 2009; Steele & Powell, 1992; Stojanowski
et al., 2013).
Whatever the shortcomings of Turner’s hypothesis of the peopling of the Americas based
on his analysis of the dental record, it nonetheless reveals the challenges for developing a
broad assessment of an exceedingly complex record involving genetic drift, gene flow and
subsequent admixture, and other factors that result in the variation seen in the recent past
and living populations. Authorities agree that Northeast Asia is the originating point of
departure for populations that are ancestral to all native peoples living in the Americas

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.3 Biohistorical issues: temporal perspectives 381

(Tamm et al., 2007; Raff et al., 2011; Rasmussen et al., 2014). Moreover, when viewed in the
broader perspective of peopling of the planet following the spread of early modern Homo
sapiens from Africa some 150 000 years ago, the migration to and settling of the Americas
happened rapidly, possibly within several hundred years of the appearance of the founding
population(s) on the eastern side of the Bering land bridge. Analysis of this complex record of
the peopling and history of native populations in the Americas is revealed through the
molecular record. That is, mtDNA sequence data in modern Native Americans indicates that
more than 90% of mtDNA variation is limited to one of four primary lineages (haplogroups
A, B, C, and D) and a fifth lineage (haplogroup X), found throughout the American
continents and intermittently in Asia (Raff et al., 2010; Schurr et al., 1990; Torroni et al.
1993). In Mongolia, Tibet, and Siberia, these haplogroups occur together (Kolman et al.,
1996), confirming the Northeast Asian origin for the native populations of the Americas.
Torroni and coworkers (1993) regarded the main four haplogroups as representing four
founding lineages. One lineage identifies populations containing an Asian-specific nine-base-
pair (9-bp) deletion between the tRNALys and Cytochrome Oxidase II genes, which has also
been identified in skeletal remains from North American archaeological samples (Kaestle,
1995, 2010; Merriwether et al., 1994, 1995; Parr et al., 1996; Stone & Stoneking, 1993).
However, the distribution in Asia of the four haplogroups suggests just one migration from
Asia (Bonatto & Salzano, 1997; Merriwhether et al., 1995). Similarly, there is a relatively
limited range of Y-chromosome variation and the ubiquitous presence of just two lineages–
C and Q – is well documented in Native American men. Furthermore, like mtDNA variation,
the commonness of the Y-chromosome lineages coupled with the high frequencies in northern
Asia reveal the certainty of an ancestral-descendant relationship between Asia and the
Americas and the likelihood of very few migrations from the former to the latter, quite
possibly just one in the range of 18 000–15 000 yBP (Mulligen et al., 2004; Zegura et al.,
2004). Moreover, this record is confirmed in analysis of autosomal DNA (e.g., STRs), suggest-
ing a single migration in the late Pleistocene (Mulligan et al., 2004).
Wallace and Torroni (1992) traced the origin of the 9-bp deletion to southeastern China,
suggesting a common ancestral group for Macro-Indians and populations from this region of
Asia. The distinctiveness of Nadene mtDNA suggests the presence of a separate and more
recent founding population (Wallace and Torroni, 1992). Much of this genetic record and
reconstructed population history of the peopling of the Western Hemisphere derives from
lineages and genetic variation in living people. The ancient DNA record, while certainly much
smaller, is providing confirmation and clarification of some important points, especially
regarding intercontinental colonizations and population movement in the later Pleistocene
(O’Rourke & Raff, 2010; Raff et al., 2011) and regional population histories (see later).
Although the record of aDNA is drawn nearly exclusively from mtDNA and only several
areas of the Americas are sampled to any extent – especially western North America and
western South America – the record is informative about population history following the last
glacial maximum. Overall, the narrow variation in mtDNA is consistent with a single major
migration event from Asia. Analysis of single nucleotide polymorphisms (SNPs; n¼364 470)
from numerous living Native American and Siberian groups, however, suggests that following
the earliest wave, there were two later migrations, including one serving as the ancestry for
speakers of Eskimo-Aleut languages and the other for Na-Dene-speaking Chipewyan from
Canada (Reich et al., 2012) (Figure 9.7). This record suggests that the vast majority of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
382 Biological distance and historical dimensions of skeletal variation

11 168

Yoruba

56 8

3 8 138 3

18 57% 51 90% 43% 10% 12

Algonquin Han

6 65 16 21 32 8%

Zapotec 1 Karitiana Aleutian Chipewyan

16 92%

8 2 13

East Greenland Inuit West Greenland Inuit Naukan

Figure 9.7 Gene flow, reconstructed through modern day single nucleotide polymorphisms,
from Asia to America suggests two migrations occurring after the initial migration into
the Americas: first migration (blue), second migration (green), and third migration (red).
Dotted lines represent mixture events, and solid circles denote hypothesized ancestral
populations. (From Reich et al., 2012; reproduced with permission of authors and
Nature Publishing Group.) A black and white version of this figure will appear in some
formats. For the color version, please refer to the plate section.

New World populations are descended from the first migration. Moreover, the growing
consensus based on genetic evidence for one early migration and dispersal throughout the
Americas is ultimately consistent with the tripartite model proposed by Turner.
A number of authorities identify two broad clusters of craniofacial morphology in the
Holocene of South America (González-José, Neves et al., 2005; Hubbe et al., 2010, 2011;
Neves & Hubbe, 2005; Neves et al., 2003, 2004, 2005, 2007) and North America (Jantz &
Owsley, 2001; Powell & Neves, 1999). In both continents, the authors argue for the presence of
an early group and a late group, with the former having distinctive craniofacial morphology
characterized by greater robusticity, longer, narrower crania, and greater midfacial prognath-
ism than the latter group. In South America, Neves and collaborators see the presence of two
separate founding populations, with the later representing the origin of “Mongoloid” variation
(Neves et al., 2003, 2004, 2005, 2007). In North America, Jantz & Owsley (2001) viewed a
similar pattern of craniofacial differences in earlier and later groups as a discontinuity, with
the earlier not serving as an ancestor for the later.
Thus, while the molecular evidence points to a single early founding population in the
Americas forming the ancestry for most native North and South American populations, the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.3 Biohistorical issues: temporal perspectives 383

craniofacial morphological evidence suggests at least two. It may be the case that the
morphological discontinuity is due to genetic drift, natural selection, or alternatively, devel-
opmental plasticity and response to changing masticatory environments and mechanical
loading patterns due to dietary shifts, or some combination of these processes (Eggars et al.,
2011; Perez et al., 2009; and see Chapter 7). In order to address these alternative explanations,
it will be necessary to document and interpret craniofacial variation in populations that are
intermediate between early (Paleoamericans) and late (modern/living native populations).
Thus far, that work has been limited in scope.

Great Basin: population replacement or in situ evolution?


Among the best-known Native American linguistic records is that from the Great Basin of
western North America, encompassing the modern state of Nevada, bounded by parts of
California, Oregon, Utah, Colorado, and Arizona. Owing to the late date of earliest Euroamerican
contact with native populations in the early nineteenth century, native languages were preserved,
thus allowing early linguists, anthropologists, and others to study and record key languages,
some of which continue to be spoken in the twenty-first century. This linguistic record
suggests that a group speaking languages from the Numic family spread into the Great Basin
via a southern route either from southwestern Nevada (Fowler, 1992; Lamb, 1958) or from the
Death Valley region of California (Steward, 1940). The timing of this spread is unknown, but
glottochronological evidence suggests that Numic languages first appeared in the Great Basin
around AD 1000. Like many other issues discussed in this chapter, questions involve whether
this spread occurred via demic or cultural diffusion – Was it people speaking a Numic language
or the spread of language via cultural means? Was there a population replacement of pre-
Numic speakers with Numic speakers? The Numic expansion hypothesis – namely, that it was
Numic speakers who moved into the Great Basin – has taken on considerable importance
because it may explain why there were clear shifts in technology occasioned by the arrival
of new people bringing with them their styles of projectile points, basketry, and pottery
(Cabana et al., 2008; Grayson, 2011; Kaestle, 1995; Kaestle & Smith, 2001; Kaestle et al.,
1999).
The study of aDNA from western Nevada (Stillwater Marsh and Pyramid Lake) offers
an opportunity to address the Numic expansion hypothesis. Like so much of the evidence,
the answers gained from the analysis of aDNA and modern DNA provide a suggestive
picture of molecular variation. In particular, aDNA shows a genetic record in prehistoric
populations that is certainly different from that of living native groups in the region
today. This genetic record suggests that modern tribal groups at first contact in the
nineteenth century derive from Numic-speaking founder groups that had migrated into
the Great Basin. In particular, the aDNA and its associated haplogroup frequencies from
western Nevada are statistically different from all modern native-language groups except
the Penutian, the proposed founding language group by proponents of the Numic expan-
sion hypothesis. Especially striking is the very high frequency of haplogroup D in the
prehistoric western Great Basin, in strong contrast to the eastern Great Basin (O’Rourke
et al., 1999). Moreover, the frequency of haplogroup B in the eastern Great Basin is
considerably higher than in the western Great Basin (Kaestle & Smith, 2001; Kaestle
et al., 1999; O’Rourke et al., 1999; Parr et al., 1996). This spatial pattern of variation
suggests the presence of two separate population histories – East and West – for this vast
area of western North America.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
384 Biological distance and historical dimensions of skeletal variation

American Southwest
The American Southwest is among the most dynamic regions of North America. Populations
living there engaged in considerable movement, demographic disruption, and subsistence
change. The remarkable adaptations to extreme environments and successes in subsistence
productivity involving maize agriculture in later prehistory provided the context for the
development of complex Puebloan societies in the region. As with many other areas of
the world, a central issue has been explaining the origins of farming. Did early farmers in
the region migrate from elsewhere, did the idea of farming spread through cultural borrowing,
or was farming the product of local in situ innovation? Biodistance analysis provides an
important perspective that is helping to reveal the histories of the populations of this region
and to increase our understanding of this important issue. A common theme in the explan-
ation of population history in the region is the debate over the origins of agricultural
Puebloan societies. Did these groups originate in Mexico, replacing earlier hunter-gatherer
Basketmaker populations? Beginning with a comparative biodistance analysis of crania from
the Zuni site of Hawikku, Seltzer (1944) questioned the long-held notion that Basketmaker
societies had been replaced by later-arriving groups from Mexico, who carried with them their
agricultural practices. His study, along with later biodistance analyses focusing on cranial and
dental variation, provides an important context for understanding population history in the
region (Bennett, 1973; Corruccini, 1972; Schillaci & Stojanowski, 2003). LeBlanc and co-
workers (2008) used dental discrete traits to test the hypothesis that farming derived from
migrating populations from Mexico. Their analysis of dental series from the North American
Desert West and Basketmaker populations in the northern Southwest reveals that the early
Mimbres farmers share little biological affinity with populations from northern Mexico,
thereby providing little evidence of continuity between the populations of northern Mexico
and the American Southwest. Such findings do not support a model involving migration of
farming groups from Mesoamerica to the American Southwest. Instead, these findings are
consistent with a growing body of molecular data, especially from living societies, that offer
little evidence that populations from Mesoamerica immigrated to the Southwest. In particular,
the distribution of mtDNA haplogroups among modern Native American populations across
the Southwest and northern and central Mexico is strongly inconsistent with a population
expansion from south to north (Malhi et al., 2002, 2003; Watson, 2010). As such, the
bioarchaeological record not only provides additional evidence for the in situ adoption and
intensification of maize farming by local indigenous foragers, but it also provides support to
the growing notion that common ancestry of modern tribes may have considerable antiquity
in the Southwest (Lorenz & Smith, 1996; Torroni et al., 1993; Watson et al., 2010).

Great Plains
The prehistoric and early historic North American Great Plains region was the scene of a
number of population movements that have implications for biological history. One of the
most important of these developments involved the movement of Caddoan speakers (Pawnee
and Arikara) northward up the Missouri Valley into the present-day states of Nebraska and
South Dakota. This movement is well documented on the basis of linguistic and archaeo-
logical evidence (Jantz, 1973; Parks, 1979), providing an important context for biodistance
analysis and identification of indigenous and immigrant groups. Building on preliminary
research by Bass (1964), questions regarding these relationships have been addressed by

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.3 Biohistorical issues: temporal perspectives 385

various researchers (Jantz, 1972, 1973, 1974, 1977, 1994; Jantz et al., 1981; Key, 1983, 1994;
Key & Jantz, 1990; Owsley & Jantz, 1978; Owsley, Morey et al., 1981; Owsley, Slutzky et al.,
1981; Owsley & Symes, 1981; Ubelaker & Jantz, 1979; and see later).
Multivariate principal components analysis of more than 800 crania from Kansas,
Nebraska, South Dakota, and North Dakota reveals a general continuity of the native
populations from the initial Paleoamerican (c. 10 000 yBP) occupation of the region to the
Woodland period (to c. AD 900) (Key, 1983). In late prehistory and into the late contact
period, local population continuities are suggested, including one linking prehistoric
Woodland groups in the northern Plains with the historic Mandan of North Dakota.
A biological discontinuity is indicated between populations representing the Central Plains
tradition (c. AD 900–1400) and the period immediately preceding it – the Woodland
period. Beginning with the Central Plains tradition and extending to the historic Arikara
period, some craniofacial changes are documented, including an increase in facial height
and a decrease in vault height. This change may represent admixture with the Mandan
(who have distinctively low cranial vaults), the indigenous population encountered by the
immigrating Arikara in the Upper Missouri Valley (Jantz, 1972, 1973, 1977; Jantz et al.,
1981). Thus, unlike craniofacial changes documented in some other North American
settings, the pattern of biological change in the northern Great Plains appears to indicate
the influx of new populations around AD 900.

American Midwest (west-central Illinois)


The long-term investigation of population history in west-central Illinois conducted by
Buikstra and coworkers has played an instrumental role in the development of method,
theory, and application of biodistance analysis in bioarchaeology. This research resulted in
an impressive data set that addresses issues relating to the identification of biological continu-
ity and discontinuity in the past (Buikstra, 1975, 1976a, 1977a, 1980; Conner, 1984, 1990;
Droessler, 1981; Konigsberg, 1987, 1990; Konigsberg & Buikstra, 1995). In her preliminary
biodistance analysis of 2000 years of prehistory encompassing the Middle Woodland (50 BC–
AD 300), Late Woodland (AD 300–800), and Mississippian periods (post-AD 800), Buikstra
asserted that migration is not the correct model for explaining human biological change in
this setting, but rather such change represents “a response to changing patterns of local
adaptation” (1975:293).
Follow-up studies lend strong support for Buikstra’s initial assessment of the biological
history of prehistoric west-central Illinois. Craniometric analysis of Late Woodland and
Mississippian period samples (Yokem, Ledders, Klunk, Schild, Koster sites) reveals a number
of significant morphological differences between periods – Mississippian crania exhibit wider
faces, smaller interorbital distances, larger zygomatics, and larger occipital condyles than Late
Woodland crania (Droessler, 1981). Two competing models best explain these temporal
changes: (1) biological change is due to influx of a new population at the close of the Late
Woodland period; or (2) biological changes are responses to circumstances taking place within
the region (Droessler, 1981). Multivariate (discriminant) statistical craniometric analysis
reveals low and statistically insignificant distance values between Late Woodland and
Mississippian samples for both males and females, leading Droessler to conclude that “the
results lend greater support to the in situ development model than to the migration model”
(1981:184).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
386 Biological distance and historical dimensions of skeletal variation

Biodistance analysis based on nonmetric cranial traits from an expanded sample encom-
passing the Middle Woodland, Late Woodland, and Mississippian (post-AD 800) periods is in
general agreement with Buikstra’s and Droessler’s assessments of population history and
biological continuity (Konigsberg, 1987). The striking consistency of temporal ordering of
samples based on biodistance analysis, along with archaeological evidence for continuity,
provides compelling evidence that the rise of complex societies in west-central Illinois was an
in situ process rather than an event involving significant external forces or replacement.
This is not to say, however, that there were neither population movements nor changes in
the patterns of gene flow within the region. Analysis of cranial morphology in the region
reveals a record of changing patterns of gene flow, especially an expansion in the range of
gene flow in the earlier Late Woodland populations relative to that seen among the later
Mississippian populations (Steadman, 2001). Gene flow is so dramatic during the Late
Woodland that biological distance is not correlated with geographic distance. By contrast,
localization of gene flow during the subsequent Mississippian period results in a classic
pattern of isolation-by-distance, such that biological distances are strongly correlated with
geographic distance. This conclusion suggests the later Mississippian populations participated
in patterns of migration and marital exchange that result in more localized gene flow than
those of their predecessors.
This finding is consistent with other analyses of degree of gene flow in the American
Midwest. In the Ohio River valley, Tatarek & Sciulli (2000) suggested that, owing to greater
mobility of foragers than farmers, inter-population biological differentiation should be less
across forager populations than in farming populations due to their greater level of gene flow
between local populations. They tested this hypothesis through a biodistance analysis of a
large series of Late Archaic (pre-1000 BC) hunter-gatherers and Late Prehistoric (AD
1000–1600) agriculturalists. Tatarek and Sciulli used R-matrix models to assess population
structure. Such models provide a robust means of identifying genetic distance, population
divergence, and the overall effects of gene flow into a population (Relethford, 1994, 2012).
Consistent with their hypothesis, the analysis yielded evidence of greater inter-population
differentiation among Late Prehistoric farmers relative to Late Archaic hunter-gatherers, who
show remarkably little differentiation on the whole. Thus, it appears that reduced mobility
among members of these later farming societies contributed to a general reduction in gene
flow between local populations.

Spanish colonization and biological interaction in the American Southeast


Some of the earliest European exploration and colonization efforts in the New World were
directed at the Atlantic seaboard, in the modern states of Florida and Georgia. Founded by
Ponce de León in 1513 and subsequently colonized a half-century later, La Florida was among
the first large-scale European efforts to exploit native populations in the Americas. Over the
next two centuries, a complex mission system was established mostly along the Atlantic
coast, resulting in alteration in lifestyle and activity due to labor exploitation (see Chapter 6)
and social disruption, leading to changing patterns of identity, biological interaction, and
population structure. Biodistance analysis of dental and cranial traits reveals patterns of
continuity and discontinuity in the region, showing, for example, links between ancestral
and descendant populations during the mission period as communities were forced southward
in conjunction with expansion of British interests (Griffin et al., 2001). On the other hand,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.3 Biohistorical issues: temporal perspectives 387

0.2

Apalachee
Eigenvector 2 (39.4%)
Guale
0.1 Coastal

Irene Mound
0

Timucua
–0.1 Irene Mortuary

–0.2
–0.2 –0.1 0 0.1 0.2
Eigenvector 1 (54.8%)
Figure 9.8 Eigenvector plot depicts the genetic differences, represented by tooth dimensions,
between precontact populations in Georgia and Florida. (From Stojanowski, 2005a;
reproduced with permission of author and the American Anthropological Association.)

analysis of dental biodistance for La Florida indicates a dynamic period of redefinition and
emergence of groups. These ethnogenetic processes resulting in new social ties between
communities and emergence of new groups in northern Florida and coastal Georgia reveal
a complex period of biological history and social interactions (Figure 9.8).
Large-scale regionally based biodistance analysis provides essential perspective on social
and cultural formation, identity, and interaction (Stojanowski, 2004, 2005a, 2005b, 2005c,
2009, 2013a, 2013b; Stojanowski et al., 2007). In this regard, La Florida reveals population
interactions within and among communities in the mission setting on this remarkably
dynamic social landscape. Overall, the process of missionization involved aggregation of
communities resulting in broad patterns of admixture, largely in contrast to the later prehis-
toric period when population boundaries were defined by local chiefdoms that controlled
regional resources. That is, the patterns of differentiation viewed elsewhere in the later
prehistoric Eastern Woodlands (Steadman, 2001; Tatarek & Sciulli, 2000), and likely for this
setting prior to European contact, underwent a rapid change as separate communities were
gathered through various colonial-driven reducción policies into mission centers during the
historic period (Blair, 2013; Hann, 1988, 1996; Larsen, 1990a; Stojanowski, 2005c). While the
strategy of population agglomeration was viewed as facilitation of proselytizing and conver-
sion to Christianity, the underlying practice was viewed as necessary for labor exploitation,
control of native societies, and adjustment to population losses taking place during the
colonial period.
Independent biodistance analyses of dental and cranial variation (discrete traits and
metrics) documented patterns of genetic distance (Griffin et al., 2001; Stojanowski, 2005a,
2005b, 2005c, 2005d, 2009) that show striking similarity in results. In particular, there is an
increase in phenotypic variability in the early mission period followed by a decline in
variability in the late mission period. The earlier trend likely reflects the aggregation of
diverse groups in a single community and thus increases in biological diversity, largely
reflecting the presence of what were originally distinct populations now gathered into a

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
388 Biological distance and historical dimensions of skeletal variation

1600–1650 1650–1704
0.2 0.2
San Martin

Eigenvector 2 0.1 0.1

Eigenvector 2
Ossuary Yamassee
0.0 0.0
SCDG-Amelia
Patale San Luis

–0.1 –0.1
SCDG-SGI

–0.2 –0.2
–0.2 –0.1 0.0 0.1 0.2 –0.2 –0.1 0.0 0.1 0.2
Eigenvector 1 Eigenvector 1
Figure 9.9 Eigenvector plots for early mission (left) and late mission (right) populations in La
Florida indicate a reduction in genetic diversity among local groups over time. (From
Stojanowski, 2009; reproduced with permission of authors and University Press of Florida.)

single population. However, in the later mission period of the later seventeenth century, a
population bottleneck and increased broad-scale gene flow among once-distinct populations
resulted in reduced genetic diversity in the larger, aggregated mission community (Figure 9.9).
Although the patterns of genetic diversity presented in these analyses are not a surprise in
consideration of the context of population aggregation and movement during the mission period,
they nonetheless reveal the powerful analytical depth of biodistance analysis for developing
a more informed understanding of biological variation and social change in this setting.

9.3.2 Linking the dead to the living


The call for identification of biological and cultural links between living native groups and
potential archaeological ancestors as a result of new laws on the repatriation of Native
American remains in the United States has created additional incentive for biodistance
analysis. Biodistance provides a powerful means of linking present and past populations in
North America or other regions where repatriation of human remains is a point of discussion.
On the other hand, it is difficult to identify relationships between living and ancient groups
owing to the remarkable level of population movement and decline of native societies after
initial contact by Europeans.
Alaska is a highly visible focal point of repatriation discussions, in large part due to the
presence of vibrant Native communities living in the region today. In order to identify links
between living and past groups, a substantive amount of dental data has been analyzed from
both contexts on Kodiak Island. Prehistoric and living populations have been studied by
biological anthropologists seeking information about ancestral and descendant relationships
(Scott, 1994). Scott (1994) analyzed dental morphological variables from prehistoric archaeo-
logical dentitions from the Uyak site, which includes a population succession from the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.4 Biohistorical issues: spatial perspectives 389

pre-Koniag (1500 BC–AD 1100) and Koniag (AD 1100–1763) periods. Univariate statistical
treatment of crown trait frequencies reveals that only two (of 14) traits exhibit significant
differences between the pre-Koniag and Koniag archaeological populations. Analysis of living
Koniag populations reveals no significant intervillage variation. Trait comparisons indicate
only three statistically significant differences between pre-Koniag and modern Koniag and
two differences between prehistoric and modern Koniag. This finding suggests long-term
population continuity.
In order to frame the biodistance analysis of native Kodiak Islanders in a wider biogeo-
graphical context, Scott (1994) compared crown trait frequencies between Kodiak Island
samples (three prehistoric, three modern) and prehistoric and protohistoric northern popula-
tions (Eskimo, Aleut, Northern Maritime, Central Maritime, Gulf of Georgia, Athapaskan,
St. Lawrence Island). Biodistance analysis based on a symmetrical relationship matrix reveals
important links between prehistoric and historic groups in the larger Arctic region, suggesting
broad patterns of continuity. Importantly, this analysis indicates that questions of local
continuity need to be addressed within larger geographical settings. With regard to questions
of affiliation and cultural patrimony, the dental record provides evidence from skeletal
remains that the prehistoric Kodiak Islanders are the ancestors – in at least some measure –
to native individuals living in the area today.

9.4 Biohistorical issues: spatial perspectives


9.4.1 Ethnic boundaries and territoriality
Archaeologists devote an enormous amount of attention to defining cultural boundaries
between broadly contemporaneous populations (Lightfoot & Martinez, 1995). In the afore-
mentioned west-central Illinois samples, boundaries between populations are defined with
some degree of precision based on stylistic variation and other differences in material culture
(Farnsworth & Asch, 1986; Farnsworth et al., 1991). In some respects, processes underlying
cultural change and cultural identity are less understood than processes underlying biological
change and population identity. Biodistance analysis based on skeletal and dental morph-
ology potentially represents a more tractable alternative to boundary definition than analysis
of artifactual remains (Konigsberg & Buikstra, 1995).
Konigsberg and Buikstra (1995) applied Womble analysis (“wombling”), a multivariate
population distance analysis that is especially useful for identifying spatial biological discon-
tinuities between populations. This type of analysis attempts to identify regions within “maps”
of biological variables. The resulting discontinuities serve to identify boundaries that may
have limited biological exchange. Womble analysis of cranial nonmetric traits from skeletons
from 10 prehistoric sites demarcates possible population boundaries. The Ray site (Middle
Woodland period), located farthest north in the Illinois River valley, is distinctive, suggesting
that it is outside the boundary that demarcates the other nine sites (Figure 9.10). The Yokem
and Elizabeth sites appear to lie across or within a boundary zone that separates a southern
from a northern series of populations. The argument for boundary identification via bio-
distance analysis is especially convincing, as the Ray site, located in the central Illinois Valley,
contains stylistically different artifacts and different site organization than the cluster of sites
associated with the lower Illinois Valley. The Elizabeth site is in the region that demarcates the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
390 Biological distance and historical dimensions of skeletal variation

iver
ois R
Missis Illin
s
ippi R

RY
iver

EZ

YO

LD KL
GI
SH
KO
HN
HR

Figure 9.10 Map generated from Womble analysis of cranial discrete traits in west-central
Illinois populations. The analysis identifies a series of tiles collectively representing
population boundaries. The map indicates that the Ray site (RY) to the north is separate
from the other sites, and the Yokem (YO) and Elizabeth (EZ) sites lie in a population
boundary separating a northern from a southern group (LD, Ledders; KL, Pete Klunk;
GI, Gibson; SH, Schild; KO, Koster Mounds; HN, Helton; HR, Hacker South Mound 2).
(From Konigsberg & Buikstra, 1995; reproduced with kind permission from Springer
Science and Business Media.)

lower from the central Illinois Valley, and contains a mix of cultural and settlement attributes
from both regions. Overall, this site appears most like lower valley archaeological sites. Thus,
biodistance analysis locates the study of population groupings within a behavioral realm that
has been based previously on conventional archaeological data.

9.4.2 Family ties: inferring kinship, relatedness, and postmarital residence


practices
Biodistance analysis provides important contexts for reconstructing and interpreting small-
scale biological relationships and drawing inferences about familial relationships and kinship
patterns. Rösing (1986) analyzed craniometric and cranial nonmetric traits in individuals from

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.4 Biohistorical issues: spatial perspectives 391

two Egyptian cemeteries, Qubbet el Hawa and Elephantine. Qubbet el Hawa is the location for
burial of nobility from the first upper Egyptian capital on the island of Elephantine. The site of
Elephantine contains the remains of mostly middle- to low-status individuals. Groups of
individuals from specific graves exhibit unusual frequencies of cranial nonmetric traits. In
grave 89, for example, 47% of individuals have occipital precondylar tubercles, normally a
very rare occurrence. This frequency is considerably higher than found in other populations
worldwide (compare with Hauser & De Stefano, 1989:135). The grave owner and lineage
founder – Sebek-Hotep – shares six of seven variants with the other individuals interred in the
family tomb. These findings provide corroboratory evidence for familial relationships and
likely a strong record of endogamy.
The study of skeletons from historic-era rural family Euroamerican cemeteries in North
America facilitates inferences about familial relationships. Analysis of remains from Wise
Cemetery in Ontario (Spence, unpublished data) and Cross Cemetery in Illinois (Larsen,
Craig et al., 1995) reveals an unusual prevalence of traits. In the Wise cemetery, the
association between grave markers and individual skeletons permitted genealogical recon-
struction of specific family relationships, thus providing important context for biodistance
analysis. Father and son (Peter Wise Sr. and Peter Wise Jr.) show evidence of trait
correspondence for some unusual variants, which include divided supraorbital foramen,
absence of zygomaticofacial foramen, open foramen spinosum, temporal squamous fora-
men, and patent mendosal suture, among other traits. The general pattern of consistency in
trait distribution in these two individuals as well as the homogeneity of traits in the sample
is well outside of what would be expected in a random sample of individuals drawn from a
larger population.
The Cross series contains an unusually high frequency of individuals with partial or full
persistent metopic sutures: nine of 14 crania (64.4%) have either partial or full metopic
sutures. This frequency contrasts sharply with what is found in other populations, most of
which are well under 10% (Hauser & De Stefano, 1989; Sullivan, 1922), with few exceptions
(Alt & Vach, 1998). Metopism heritability is unknown, but most workers are confident that it
has a high genetic component (Hauser & De Stefano, 1989). Unlike the Wise series, precise
familial relationships among skeletal individuals could not be determined owing to the lack of
association between grave markers and human remains. Historical records indicate that
individuals interred in the cemetery included primarily members of a nuclear family and their
close kin (Larsen, Craig et al., 1995). Thus, an extraordinarily high frequency of metopism is
consistent with archaeological and historical documentation that the cemetery is dominated
by closely related individuals.
There is a striking degree of homogeneity of traits in skeletal samples from family
cemeteries. When observed in large skeletal samples, such homogeneity suggests the
presence of endogamous mating patterns. Biodistance analysis of dental nonmetric traits
of some 300 individuals from the Early Bronze Age site, Bab edh-Dhra’, in southern Jordan
reveals an unusually high degree of homogeneity – about 80% of the sample shares specific
traits (Bentley, 1991). There is also a high degree of similarity between adult females and
males. Analysis of trait distribution reveals a clustering of rare traits within particular
tombs. For example, only 8% (13/158) of the Bab edh-Dhra’ population possess mandibular
molars with six cusps; all individuals with the trait are clustered in six of 25 tombs. Third
molar agenesis is present in only five tombs. In summary, the high overall degree of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
392 Biological distance and historical dimensions of skeletal variation

homogeneity suggests group endogamy, but with the clustering of traits indicating burial of
related individuals in the same tomb.
The conclusion that trait homogeneity represents group endogamy is based on the assump-
tion that traits become homogeneous through time in biological lineages (Kennedy, 1981;
Konigsberg, 1987; Konigsberg & Buikstra, 1995). Hypothetically, adult males and females in
endogamous societies should exhibit the same degree of intra-sex variability. In contrast,
exogamous populations should exhibit a relatively high degree of variability as well as
significant sex differences. This suggests, therefore, that the natal group or non-migratory
component of the population should be relatively more homogeneous in trait expression than
the migratory component.
The remarkable transformation of mortuary behavior of native populations in regions of the
New World controlled by Spain, especially in those areas having strong missionization efforts,
is revealed not only in burial posture but also in the organization in family units. In this regard,
at Tipu, Belize, and other Spanish colonial cemeteries in the Maya region, family-owned
cemeteries were used (Jacobi, 1997, 2000). In this setting, there are clear tendencies for the
presence of clustering of dental discrete traits in specific regions of the Tipu cemetery,
identifying family or kin groups. For example, there are clusters of rare traits that identify
kin relationships, such as the presence of a labial groove on the maxillary incisors and canines
and winged central incisors. At San Luis de Apalache, a group of three elite burials located
adjacent to the altar, also the only individuals interred in coffins, had similar dental size–shape
profiles suggesting familial relationships (Stojanowski et al., 2007). No other biodistance
patterns are present in the cemetery series, suggesting that only the elite were interred in a
distinctive cluster based on kinship. At the San Pedro y San Pablo de Patale mission, Jones and
coworkers (Jones, Hann et al., 1991; Jones, Storey et al., 1991) hypothesized that interment in
the cemetery was structured according to kin identity. To test the hypothesis, Stojanowski
(2005d) undertook a multivariate biodistance analysis of buccolingual and mesiodistal dimen-
sions of permanent teeth in order to assess size and shape. In contrast to San Luis, a number of
statistically significant intra-cemetery patterns are consistent with the hypothesis of kin
patterning. For example, individuals buried on the right side of the cemetery had statistically
larger teeth than those on the left side of the cemetery. Overall, the statistical analysis supports
the hypothesis that rows were occupied by distinctive kin/family groups, a pattern that is
especially clear in relation to the location of juvenile interments (Figure 9.11).
Postmarital residence is a central area of interest in anthropology, largely due to the
importance of understanding the general organization of human societies and the establish-
ment of kin-based networks, within and between communities and regions. Moreover, post-
marital residence is a fundamental part of social, political, and economic organization (Hubbe
et al., 2009). Therefore, to understand postmarital residence is to understand some very basic
premises of how past societies were organized and how they interacted with surrounding
communities and regions (Hubbe et al., 2009; Stojanowski & Schillaci, 2006). Moreover, the
case can be made that biological data that generate sex-specific patterning in phenotypic
variation are likely more appropriate for the documentation of postmarital residence in
prehistoric contexts because other lines of evidence – such as lineal patterns of material
culture transmission (e.g., ceramic styles) – are not determined by factors that are necessarily
predictable. In contrast, biological variation is governed by principles of genetic inheritance
(Schillaci & Stojanowski, 2003; Stojanowski & Schillaci, 2006).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.4 Biohistorical issues: spatial perspectives 393

0.5

63
52
54 7
7
8
0.0 44 8 5
8 8
7
56A
7
PC 2

6 42 8

–0.5
5

–1.0
–3 –2 –1 0 1
PC 1
Figure 9.11 At the mission cemetery of San Pedro y San Pablo de Patale, Florida,
principal components – which explain tooth size and morphological variation between
individuals – indicate definitive similarities between subadult individuals in the same burial
rows and sides. Open circles represent individuals buried on the left side of the cemetery
and filled circles subadults buried on the right side. Data points correspond to burial
rows (e.g., rows 7 and 8). (From Stojanowski 2005d. Reproduced with permission of the
author. © 2005 by the Southeastern Archaeological Conference.)

In his analysis of cranial variation in the late prehistoric Irene Mound site, Georgia skeletal
series, Hulse (1941) observed less craniometric variation in adult females than males, leading
him to speculate that this Mississippian society had a matrilocal postmarital residence pattern.
While Hulse’s analysis was insightful, the idea of postmarital residence and the ability to
document it via bioarchaeological research was undeveloped until the 1970s, beginning with
the influential studies of Corruccini (1972), Lane and Sublett (1972), and Spence (1974a,
1974b). Further inferences have been drawn about postmarital residence patterns by various
researchers, especially in relation to inter-group and intra-group variance in nonmetric traits
(Birkby, 1982; Buikstra, 1980; Cook & Aubry, 2014; Corruccini, 1972; Droessler, 1981;
Kennedy, 1981; Konigsberg, 1988; Nystrom & Malcom, 2010; Schillaci & Stojanowski,
2002, 2003; Tomkzak & Powell, 2003; see review by Stojanowski & Schillaci, 2006).
Spence (1974a, 1974b) emphasized within-group variation based on his modification of a
simple matching coefficient. In his model, a high degree of similarity in traits within a group
represents low variability. Thus, a higher mean similarity coefficient (i.e., decreased variabil-
ity) in adult males of a population in comparison with adult females indicates that males are
more closely related to one another than are females. This pattern suggests co-residence of
related men – patrilocal or virilocal residence. In this marriage system, males remain in the
natal group, whereas female partners immigrate from elsewhere. Because females are drawn
from a greater number of population groupings, they display greater trait variability
than males.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
394 Biological distance and historical dimensions of skeletal variation

Spence’s biodistance analysis of cranial, dental, and postcranial traits in skeletal remains
from the Classic period urban center of Teotihuacan in central Mexico reveals a pattern of less
variability between males than between females (see Chapter 8). Moreover, some trait fre-
quencies are considerably higher in males than in females for specific apartment complexes.
At the La Ventilla B apartment compound, for example, two supraorbital traits – supraorbital
foramen presence and multiple supraorbital foramina – are far more common in males than in
females (see discussion in Spence, 1994). The greater trait similarity among males than among
females suggests that the corporate social group may have been organized around male kin
who maintained their residence over the period of the occupation of the apartment compound.
In this patrilocal residence pattern, adult females would have originated elsewhere, either
from other apartment compounds or outside of Teotihuacan altogether.
Owing to the availability of their rich ethnographic, ethnohistorical, and archaeological
records, a number of regions have developed important and compelling analyses of post-
marital residence-based skeletal and dental parameters. In the American Southwest, for
example, at the ancestral Zuni settlement of Hawikku (New Mexico), dental biodistance (cluster)
analysis reveals individual cemeteries surrounding the community display distinctive pat-
terns, especially in Cemetery 3 where two-thirds of the sample are related (Howell & Kintigh,
1996; and see Birkby, 1982; Corruccini, 1972). Building on this biodistance analysis, Schillaci
and Stojanowski (2002, 2003, 2005; Stojanowski & Schillaci, 2006) have undertaken a
comprehensive craniometric biodistance analysis for a range of late prehistoric communities
in northern New Mexico, addressing the general hypothesis that the matrilocal residence
pattern that is ubiquitous among the modern Puebloan communities today has its roots in
their prehistoric Anasazi ancestors. An R-matrix biodistance analysis reveals a complex
pattern of residence in later prehistory. Consistent with their hypothesis, at least one prehis-
toric Tewa Pueblo – Puye – shows greater male than female variability, indicating the strong
likelihood that the community postmarital residence was matrilocal (Schillaci & Stojanowski,
2005; and see Corruccini, 1972). On the other hand, biodistance analysis of Chacoan culture
Pueblo Bonito reveals two contemporaneous but distinct populations (Figure 9.12).
Biodistance analysis reveals greater female than male variability, reflecting bilocal or non-
subscribed postmarital residence (Schillaci & Stojanowski, 2003). The latter finding contra-
dicts inferences about matrilocal postmarital residence for the Anasazi (Eggan, 1950) and
especially for Pueblo Bonito (Peregrine, 2001). Regardless of the conclusions drawn from the
biodistance analyses, the skeletal record reveals that postmarital residence in the prehistoric
American Southwest was likely considerably more varied than in the native societies in the
region today.
The analysis of relatedness, including postmarital residence, at the household or neighbor-
hood level is only rarely possible, largely because of the predilection for cemetery burial,
where it is far from clear how the cemetery is structured in relation to families or kin-based
groups. In rare instances, analysis at the household level is possible. At the large Neolithic
community of Çatalhöyük, Turkey (7400–6000 BC), the deceased were interred under the
floors of houses, presumably the place of habitation at the time of death (Hodder, 2006).
Pilloud & Larsen (2011) tested the hypotheses that (1) individuals interred within the same
house were biologically related, and (2) neighborhoods, demarcated by clusters of contem-
porary houses, were also defined by biological affinity. Dental biodistance analysis (univariate
and multivariate) revealed remarkably minimal phenotypic patterning in dental size or
morphology for deciduous or permanent teeth either at the household or neighborhood levels.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.4 Biohistorical issues: spatial perspectives 395

West Cemetery
North Cemetery

Stress < 1%

Figure 9.12 Approximate location of adult burials in the West and North Cemeteries at
Pueblo Bonito, New Mexico. Linear discriminant functions, using nine craniometric traits,
classified individuals into West (open shapes) or North (shaded shapes) cemeteries. Two
adult females, represented by stars, were misclassified into cemeteries based on
craniometric traits. This evidence suggests that exogamy was practiced between these two
kin-groups. In this figure, triangles represent adult males, and circles represent adult
females. (From Stojanowski & Schillaci, 2006; reproduced with permission of authors and
John Wiley & Sons, Inc.)

There are some relatively rare traits in the Çatalhöyük series that are above the mean for
individual houses (e.g., 50% parastyle in one house). These findings suggest that biological
affinity played only a limited role in who is or is not associated with specific houses and
neighborhoods. Thus, habitation at Çatalhöyük may have been constituted by “practical kin”
rather than “official (biological) kin” (sensu Bourdieu, 1977). That is, affiliation may have
been determined not on biological terms but rather as groups were called together for various
social or cultural reasons. At this early farming community, this arrangement was likely

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
396 Biological distance and historical dimensions of skeletal variation

motivated by the adoption of domesticated plants and animals. Individuals may have aligned
themselves with practical kin for economic-related activities such as herding animals,
planting crops, harvesting, and other functions that required groups of individuals. This
interpretation of a very early farming society still highly dependent on nondomesticated
plants and animals is consistent with an analysis of co-residence patterns for a large sample
of present-day foraging societies (Hill et al., 2011). Most individuals in the sample of
residential groups are not genetically related. This finding is in sharp contrast to the long-
held assumption that hunter-gatherer residence patterns are entirely kin based (Service, 1962).
Importantly, it underscores the significance of residence for cooperation and social learning,
which are not necessarily functions that have to be carried out within biological kin groups.
As with foraging societies, biological kinship in at least this community of early farmers was
likely not the sole defining principle of social organization. This lack of phenotypic patterning
as it pertains to identification of kin in Çatalhöyük households and neighborhoods is not to
say that no other phenotypic patterning was present in this community. That is, dental metrics
of the males are considerably less variable than those of the females in this setting (Hillson
et al., 2013; Pilloud & Larsen, 2011), suggesting that for the community as a whole, men had
greater biological affinity to one another than did women. These results are consistent with a
community that practiced patrilocal residence, an important finding that counters earlier
interpretations of a matrilocal residence pattern and a matriarchal society as inferred from the
symbolic assemblages at the site, such as the so-called mother-goddess figures (Gimbutas,
1982). Thus, females may have been moving into the community on a greater scale than males
and perhaps from some distance.
While dental and skeletal variation provides an important window onto reconstruction
of postmarital residence pattern, new insights are becoming available via analysis of
aDNA, especially with the focus on variation in mtDNA. With DNA recovered from
archaeological human remains, it has become possible to evaluate genetic relationships
more directly and to draw conclusions about postmarital residence. Because mtDNA is
inherited from the mother only, its analysis provides an important perspective on the
degree of matrilineal relatedness within populations and between individuals. The recovery
of mtDNA from a number of archaeological settings allows bioarchaeologists to test
hypotheses about postmarital residence in general and matrilineal relatedness in particu-
lar. Earlier research on Middle Woodland (Hopewell) populations in Illinois, based on
skeletal biodistance analysis, revealed greater intra-site variation in females than males,
making the case that postmarital residence was primarily patrilocal (Buikstra, 1980;
Konigsberg & Buikstra, 1995). In order to test this hypothesis about Hopewell postmarital
residence, Bolnick & Smith (2007) analyzed mtDNA collected from 39 individuals from the
Pete Klunk mound group (Illinois). This analysis revealed the presence of all five Native
American haplogroups, and a population dominated by haplogroup C (n¼19). Comparison
of female and male diversity reveals considerably greater haplogroup variation in males.
While some settings suggest patterns of placement of related individuals in cemetery
contexts (see earlier), there were no spatial patterns of mtDNA lineages (Figure 9.13),
indicating that although the postmarital residence patterns were likely dominated by
matrilocal residence, matrilocal relationships were not expressed in mortuary practices.
This work underscores the importance of mtDNA in identifying social relationships in
past societies.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.4 Biohistorical issues: spatial perspectives 397

Haplogroup A
Haplogroup B
Haplogroup C
Haplogroup D

Feature A

Feature B

Figure 9.13 Spatial distribution of haplogroups in Mound 6 from the Pete Klunk mound
group, Illinois River valley. Of the 19 haplogroup C individuals in the total sample, eight are
represented here in this mound. There is no clear spatial delineation between individuals of
different haplogroups. (Adapted from Bolnick & Smith, 2007; reproduced with permission
of authors and the Society for American Archaeology.)

9.4.3 Social groupings of biological diversity


Status and rank in organizationally complex prehistoric societies have been abundantly
identified on the basis of archaeological mortuary evidence (various authors in Beck, 1995;
Brown, 1971; Chapman et al., 1981). This chapter has made clear that osteological data
provide an independent source of information for testing hypotheses about past social
systems. Toward this effort, Wilkinson and Norelli (1981) evaluated social organization and
alternative hypotheses regarding the growth and development of Monte Albán, Oaxaca, a
massive and complex archaeological site occupied by a ranked society from c. 500 BC–AD 650.
The reasons for the emergence of Monte Albán as a city and a key player in Mesoamerican
complex societies are hotly debated. Blanton (1978) argued that its rise was facilitated by
peaceful and voluntary associations of local leaders who selected the location of the site as an
administrative center because of its political and economic neutrality. Santley (1980) con-
tended that the site arose due to increasing population pressure. Dependence on limited land
and water resources led to differences in wealth between leaders in the surrounding Valley of
Oaxaca, whereby the most powerful leaders and their kin had the highest positions of
authority. These alternative models have implications for the nature of intermarriage between
classes – exogamy or interclass marriage (Blanton’s model) versus endogamy or intraclass

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
398 Biological distance and historical dimensions of skeletal variation

marriage (Santley’s model). If social classes were mostly endogamous, as Santley contends,
then biodistance analysis should reveal clear distinctions in trait pattern and frequency
between higher- and lower-rank groups. These differences would reflect restrictions on
inter-group migration and mating.
Univariate statistical comparison of dental and craniofacial nonmetric traits from tomb, or
high-status, and burial, or lower-status, contexts at Monte Albán reveals no significant
differences between the two subgroups (Wilkinson & Norelli, 1981). This finding is consistent
with the model that the two subgroups are drawn from the same larger population. Most
importantly, consistent with Blanton’s hypothesis regarding relatively open interclass com-
munication, these results suggest that Monte Albán class structure may not have been
strongly endogamous.
Other social distinctions aside from rank have also been explored via biodistance analysis.
Multivariate craniometric analysis of historic-era Arikara from the Leavenworth site, South
Dakota, reveals a very high degree of biological variability, far exceeding what would be
expected in a population practicing village endogamy (Byrd & Jantz, 1994; Key & Jantz,
1990). Comparisons of this series with populations that were known to have been distinctive
in earlier Arikara history (e.g., Larson site) reveal an unusually high degree of heterogeneity in
craniometric variation. Consistent with archaeological evidence for use of specific burial areas
by groups attempting to maintain corporate identity (O’Shea, 1984) and linguistic evidence
for the presence of two dialects spoken by Leavenworth inhabitants (Byrd & Jantz, 1994),
there is a significant burial area effect in the analysis, specifically identifying two distinctive
groups from two respective areas of the site. This bipartite grouping pattern likely reflects
social subdivisions such as bands or attempts at maintaining earlier village identities. During
the late eighteenth century, different bands with their own social, historical, and linguistic
identities amalgamated into larger groupings, mostly in response to disease, depopulation,
and warfare. This biodistance analysis accounts for the large degree of heterogeneity in the
sample as a whole by identifying those patterns that contribute to variation. This
approach serves as an important means of revealing patterns of intra-population variability
(Raemsch, 1995).
The northern Peruvian coast provides an ideal natural laboratory for social reconstruction
in general. Among the best known cultures in the region is the Sicán (or Classic Lambayeque)
culture, lasting from c. AD 900 to 1100. During the fluorescence of the culture around
AD 1000, a considerable region of northern Peru was controlled or influenced by Sicán. At
Sicán, Huaca Loro, one of six monumental mounds, was constructed covering dozens of
massive shaft tombs, two of which have been systematically studied. They include the East
Tomb containing five individuals and the West Tomb containing 24 individuals (Corruccini &
Shimada 2002; Shimada et al., 2004, 2005). In addition, other burial features at the site –
including a “North Trench” containing five individuals with very few grave goods – may
represent a non-elite sector of Sicán society. The organization of burials and their treatment
involve careful planning and nonrandom placement of individuals within socially relevant
contexts, including a central chamber containing an elite adult male with an abundance of
material culture reflecting his high status in Sicán society, and adjacent niches with females
who may have been sacrificed. To the north and south of the central tomb are two opposing
groups, each containing nine females. The remains of these females were disarticulated and
hence they may represent individuals who were moved from an adjacent region of the West

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.4 Biohistorical issues: spatial perspectives 399

tomb. A number of questions emerge from this setting where the mortuary architecture and
arrangement of interments was planned: Was this an elite family mausoleum? Are the females
more closely related than the males, suggesting a matrilocal, matrilineal society that formed
an essential framework of this early state society? Is there a pattern of variation suggesting
that the two groups of females derive from a different level of society?
In order to address these questions, Shimada and coworkers undertook a comprehensive
multivariate biodistance analysis based on large dental series. Several distinctive biologically
informed, socially relevant patterns emerge from the analysis, including (1) a high degree of
biological similarity within the south females and northern trench individuals; (2) strong
dissimilarity with all other interments in the site, a pattern based on the sharing of canine
tubercles, maxillary distal premolar paracone expansion, strong development of maxillary
incisor shoveling, and large mesial-to-central fovea maxillary central molar chords; and (3)
dissimilarity among the north females. The basic pattern shows the division of social grouping
of some manner within a carefully planned elite mortuary context.
Although it is not possible to identify the nature of the relationships in this society, it is
clear that the inclusion of specific individuals, both elite and non-elite, was an important
characteristic of the West tomb, carefully planned in all respects. As the dental biodistance
analysis reveals, the two groups of women may reflect different lineages altogether. In order
to test this hypothesis, Shimada and coworkers (2005) undertook a potentially more informa-
tive mtDNA analysis in order to identify degree of relatedness and population origin. Consist-
ent with much of the dental biodistance analysis, the mtDNA analysis revealed a high degree
of affinity in haplogroups, especially among South females. This finding underscores kin
affiliation within each group but not across groups. That is, unlike the Hopewell populations
from North America, the Sicán mtDNA analysis revealed spatial patterning of mtDNA,
especially involving the placement of possibly two maternal kin relationships (Figure 9.14).

9.4.4 Individuals in foreign territory


In archaeological settings where contacts between disparate biological groups were known to
have occurred, biodistance may reveal the presence of biologically (genetically) distinctive
individuals drawn from different social or geographic contexts (see Chapter 8 for discussion
of migration and residence based on stable isotope analysis). Two such settings have been
explored in some detail regarding contact between Europeans and native New World groups,
namely historic-era Easter Island in the eastern Pacific (Owsley et al., 1994) and the northern
Great Plains (Jantz & Owsley, 1994b).
European contact with Easter Islanders initially occurred in the early eighteenth century.
From the outset, Europeans perceived the native people as the primary resource for exploit-
ation. In addition to providing an important labor source, the island was known to sailors “for
the charms of its women” (Owsley et al., 1994:163). Sexual relations between Europeans and
native islanders are well documented in historical accounts. Multivariate (discriminant func-
tion) analysis of cranial measurements from Easter Island populations reveals strong evidence
for presence of Europeans among the native individuals (Owsley et al., 1994). Visual inspec-
tion of three individuals shows distinctive traits that are characteristic of European and
European-descent populations. For example, a young adult male cranium from the Hekii site
has features – large and straight nasal bones, parabolic palate, well developed nasal spine, and

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
400 Biological distance and historical dimensions of skeletal variation

SOUTH GROUP
ANTECHAMBER
B15 B5
B8

B7
B6

NORTH NICHE

SOUTH NICHE
B21 CENTRAL
B16 B24 CHAMBER B20
B17

B18
B23 B10 B11
B4 PRINCIPAL B12
B9
B22 PERSONAGE
B19
B1

NORTH GROUP

B13 B14
JUVENILE
B3
B2

0 1 2
B = BURIAL N
m

Figure 9.14 Ancient mitochondrial DNA analyses of skeletons from the West Tomb in Sicán,
Peru, show the close spatial distribution of maternally related individuals in both north and
south group burials. Black shapes represent mtDNA-related individuals within the tomb,
and open circles represent individuals who are not related mitochondrially to other
individuals within the tomb. (From Shimada et al., 2004; reproduced with permission of
authors and University of Chicago Press.)

wide cranial base – that contrast with native Easter Islanders. Additionally, this and two other
individuals possess anatomical features – such as forward projection of the midfacial region –
suggesting European admixture. Owsley and coworkers contend that admixture may have
involved a European father and native mother, a conclusion well supported from the historic
accounts of native and non-native interactions.
In the Great Plains, European exploration and settlement increased the opportunity for
admixture between Europeans and native groups. There are a number of historical accounts of
Euroamerican resident traders who married native women and raised offspring (Jantz &
Owsley, 1994b). Multivariate analysis of human remains from Swan Creek, South Dakota,
indicates the presence of a morphologically unique cranium – an old adult male from a
multiple grave including five individuals. The cranium possesses a number of European or
Euroamerican traits, including reduced facial prognathism, triangular parabolic palate, prom-
inent nasal bones, and narrow nasal aperture. Three craniofacial indices that discriminate
between population groups suggest that the outlier male has Euroamerican ancestry (Gill &
Gilbert, 1990). Based on the archaeological and mortuary context, the individual was likely a
resident trader or a captive. The “foreign” origin of the individual is confirmed by his unique
element composition of dental enamel (Schneider & Blakeslee, 1990), suggesting that the
individual’s enamel had formed while he was living in another geographical setting. The

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
9.5 Summary and conclusions 401

similarity of tooth wear pattern with other individuals in the Swan Creek series suggests that
the individual lived a predominantly native lifestyle, at least with respect to the use of the
dentition in masticatory functions. The teeth are heavily worn, a pattern typical of native but
not Euroamerican dentitions in this setting.

9.5 Summary and conclusions


Biodistance analysis based on skeletal and dental morphology and aDNA identifies patterns of
biological relatedness between and within populations. Contrary to the assertions of some
(Schindler, 1985; Schindler et al., 1981), biodistance analysis is not simply a modern attempt
at racial typology. Rather, contemporary approaches transcend racial identification and
simplistic approaches to population history. These studies reveal continuities and discontinu-
ities that are valuable for showing how past populations were structured and for interpreting
key biological trends when viewed in a temporal or spatial perspective, including disease
history, activity patterns, dietary change, and other parameters discussed in this book. Intra-
population analysis also serves to identify important social parameters, such as inter- and
intra-class contacts, marriage practices, familial or kin groups, and postmarital residence
patterns. The importance of biodistance analysis is underscored in interpreting variation
between subgroups of individual populations, such as differences in stature between high-
and low- status adults. Importantly, the practical application of biodistance analysis has been
highlighted in cases involving repatriation and the identification of links between living
native populations and their potential ancestors.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:40:37, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.011
Bioarchaeological paleodemography:
10 interpreting age-at-death structures

10.1 Introduction
There has been a long and enduring interest in the demographic structure of human popula-
tions, past and present. Thomas Malthus (1798), founder of the science of demography, linked
mortality, survival, and success of a population to the food supply – populations with adequate
nutrition and lower disease loads will grow, but malnutrition and disease will limit their size.
The dynamic interplay between fertility, mortality, and available food resources formed a key
element of Charles Darwin’s (1859) theory of natural selection. Darwin concluded that
individuals having characteristics that promote their ability to survive, such as the successful
acquisition of food and resources generally, will live to adulthood and reproduce. The various
demographic measures of the success (or failure) of human populations – mortality, fertility,
survivorship, longevity, and life expectancy – have endured as important foci of inquiry by
social, behavioral, biological, and medical scientists into the twenty-first century and will
continue to do so for the foreseeable future.
Although fertility, mortality, and other key parameters of demographic analysis are import-
ant measures of success in today’s world, life expectancy remains the gold standard for
evaluating health and well-being around the globe, especially in comparative perspective.
Life expectancy of 79 years in the United States versus just 58 years in Rwanda (United
Nations, 2011) speaks to the chasm that exists in health and well-being between these two
nations, implying a considerably healthier living environment in the former than in the latter.
Similarly, increased life expectancy is an indicator of improving living conditions. In the
United States, for example, life expectancy increased from 50 years in 1910 to nearly 80 years
in 2010, clearly due, in part, to expansion in the availability of healthcare and improved
nutrition. If life expectancy (and other demographic parameters) can be measured in ancient
populations, then their age structures can provide a powerful perspective on health and
conditions of life.
Age-at-death of groups of skeletons representing past populations provides important
information regarding the demographic structure of ancient populations. As such, age-at-
death of skeletal samples is a powerful structuring phenomenon. That is, the composite age
structures of collections of skeletons are important for understanding many of the patterns of
variation discussed in the previous chapters of this book. For example, on a very simple level,
it stands to reason that a skeletal series of older individuals should be expected to have more
osteoarthritis and more dental disease than a series of younger individuals; the pathological
conditions are cumulative and are tightly linked to age. More broadly, formal analysis of age
structure of a series of skeletons derived from a once-living community or collection of
communities provides a fundamental record of well-being, as in living populations in the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
10.1 Introduction 403

St. Estève le Pont


100 Observance
Spitalfields
% Well-Preserved Bones
80

60

40

20

0
an ull

Sa ibs

et
U i
ae

co s

Fi ae
i

e
C um

Pa ora

Ti e
m
rte le

ap s
St rae

H ae

Fe ae
i
er
ad

ss nd
Sc icle

la
lla
Ve dib

Fe
nu
Sk

ln

bi
ul
um

x
R

bu
cr

m
b

te
O Ha
er

v
la

a
M

Figure 10.1 Variable frequencies of well-preserved bone elements among St. Estève le Pont
(Berre l’Etang, Bouche-du-Rhône, France), Observance (Marseilles, France), and Spitalfields
(London, United Kingdom) assemblages. This pattern reveals the disproportionate
preservation of smaller, more fragile elements. (From Bello et al., 2006; reproduced with
permission of authors and John Wiley & Sons, Inc.)

world today. For purposes of this chapter, I use Robert Hoppa’s definition of paleodemogra-
phy, “. . .the field of inquiry that attempts to identify demographic parameters from past
populations derived from archaeological contexts” (2002:9). While hypotheses and questions
in this area are dominated by bioarchaeological concerns and its primary focus is on human
remains, paleodemography explores population distributions and densities on the landscape
at any one time as well as over time from a variety of non-skeletal archaeological data sets
and documentary sources (Chamberlain, 2006; Meindl & Russell, 1998; Milner et al., 2008).
This chapter focuses on the interpretation of age-at-death profiles constructed from collec-
tions of skeletons and in relation to their archaeological contexts.
For a variety of reasons, bioarchaeological paleodemography has struggled, in both its
theory and its method, thus partly reflecting the considerable amount of debate among
paleodemographers (Buikstra & Konigsberg, 1985; Frankenberg & Konigsberg, 2006; Hoppa,
2002; Milner et al., 2008; and following). One of the most obvious issues is the relative
preservation of skeletal elements, especially smaller, more fragile elements (Bello et al., 2006;
Stojanowski et al., 2002; Walker and Johnson, 1988) (Figure 10.1). However, as discussed in
this chapter, even under the best of preservational circumstances, other more fundamental
issues abound in how bioarchaeologists interpret age structures. Despite the hand-wringing
and sometimes raucous disagreements that are alluded to in this chapter, when key assump-
tions are addressed and limitations are understood, this area of inquiry gains considerable
importance for understanding past population dynamics. Paleodemography is here to stay, for
whenever estimates of age or sex are made for skeletons from archaeological settings, a

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
404 Bioarchaeological paleodemography: interpreting age-at-death structures

demographic statement has been made. As mentioned earlier, however, this chapter makes the
case that paleodemography as it pertains to the study of human remains from archaeological
contexts is far more than estimating age and sex. Rather, the case is made that to understand
age structure is to have a more informed understanding of health, disease, and a multitude of
other factors considered in this book.

10.2 Reconstructing and interpreting age-at-death profiles: it has been


mostly about mortality
In comparison with many other areas discussed in this book, the emergence of paleodemo-
graphy as a focused area of inquiry in bioarchaeology is quite recent. Whereas analysis of
skeletal morphology and pathology and their application to the study of archaeological
skeletal series began in the eighteenth and nineteenth centuries, paleodemography as it
pertains to age structure in the past has not had a significant presence until the twentieth
century. Earnest A. Hooton’s interest in demographic structure and the publication of his
remarkable monograph, The Indians of Pecos Pueblo: A Study of Their Skeletal Remains,
marks the beginning of formal demographic analysis of archaeological skeletal series. In
particular, Hooton (1930) applied demographic methods for estimating population size and
analyzing temporal changes in population structure in order to draw conclusions about
growth and decline in a Native American population. In order to estimate death rates, he
compared the distribution of ages in the postcontact (late Glaze) period for the Pecos series,
finding that it closely approximated the profile for France in 1866–1877, giving an estimate of
death rates as 2.46/100 (1930:333). His analysis did not include some key elements of
demographic theory available at the time of his study. Nonetheless, the work provided a
thoughtful reconstruction and interpretation of age structure, certainly well in advance of
other bioarchaeological scholarship at the time and for the next several decades. Moreover,
although the Pecos Pueblo demographic data generated by Hooton are not of the quality of
later studies of this series (see discussions by Goldstein, 1953; Mobley, 1980; Morgan, 2010;
Palkovich, 1983; Ruff, 1981; Weisensee & Jantz, 2011), it set the stage for paleodemography.
Hooton is credited with the founding of paleodemography. However, it was J. Lawrence
Angel’s interest in longevity and life expectancy in the eastern Mediterranean basin (Angel,
1947, 1951) – and his inchoate application of modern theory to the subsequent study of
bioarchaeological data sets (Angel, 1968, 1969, 1971a, 1971b, 1972, 1975, 1984) – that drew
wider attention to the field, setting the stage for the later proliferation of research by
bioarchaeologists. Especially important about Angel’s work were his interests in and contri-
butions to the study of age structure for (1) identification of patterns of fertility, mortality,
survivorship, longevity, and other demographic indicators; (2) inferences regarding how
environment, social and economic conditions, and behavioral patterns have impacted quality
of life and adaptive success. Owing in large part to his influence, paleodemography became a
standard element of bioarchaeological inquiry, with respect both to stand-alone research on
age structure and as a central part of skeletal analysis. Angel’s work was at the forefront of a
group of studies in the 1960s and 1970s on past age structures (Acsádi & Nemeskéri, 1970;
Blakely, 1971; Howells, 1960; Johnston & Snow, 1961; Lovejoy et al., 1977; Ubelaker, 1974;
Vallois, 1960; and see discussions in Chamberlain, 2006; Frankenberg & Konigsberg, 2006;
Hoppa, 2002; Milner et al., 2008; Swedlund & Armelagos, 1976).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
10.2 Reconstructing and interpreting age-at-death profiles: it has been mostly about mortality 405

Fueling much of this interest in paleodemography in the 1960s and 1970s was the notion
that age-at-death composition and structure of living populations are a record of mortality
and life expectancy. That is, the mean age-at-death for an archaeological skeletal assemblage
provides a record of when, on average, individuals of a population die and the age to
which a person in the population can expect to live from a given point in their life history
(e.g., birth). One of the significant attempts to characterize age structure was Johnston and
Snow’s (1961) analysis of the Archaic period Indian Knoll series, Kentucky. At the time of
their study, the Indian Knoll series was one of the best documented and largest series of
archaeological skeletons in North America, ideally suited for bioarchaeological and paleo-
pathological study (Larsen, 2006, 2012). Using various indicators of age estimation to derive
composite individual ages for this extraordinarily large series (n¼873), Johnston and Snow
found that the average age-at-death (a proxy for “life expectancy”) in the Indian Knoll series
was just 18.6 years. Moreover, 20% of the Indian Knoll individuals were less than one year of
age, which is a considerably higher frequency than any age group in the Pecos Pueblo series.
Based on Johnston and Snow’s demographic assessment and in comparison with other
archaeological series from Austria and Egypt, they concluded that the population from Indian
Knoll had remarkably high infant mortality, “. . .similar to the more primitive, nonagricultural
groups. . .(and) little over-all similarity. . .to more economically and technologically advanced
groups” (Johnston & Snow, 1961:243).
Robert Blakely’s (1971) “mortality” analysis of Indian Knoll a decade later involving the
comparison with other Archaic period foragers and post-Archaic agriculturalists (Hopewell,
Mississippian) from Illinois revealed that, like Johnston and Snow’s analysis, the earlier
Archaic series from both Kentucky and Illinois had a higher percentage of children aged zero
to nine years old than in the later series. These earlier researchers concluded that the samples
they studied represented populations having an elevated level of infant mortality and low life
expectancy, a point to which this chapter later returns.
In ancient Greece, Angel (1947) documented a pattern of temporal increase in the percent-
age of older adults from c. 3500 BC to AD 1300, indicating to him an increase in life
expectancy based on these “mortality” profiles. In other words, like Johnston & Snow
(1961) and Blakely (1971), to Angel, higher average age-at-death or presence of a relatively
greater number of older adults translated directly into a picture of greater survivorship, length
of life, and reduced mortality overall.
Similarly, in Neolithic Çatalhöyük, Turkey (7400–6000 BC), based on his study of
remains recovered by James Mellaart, Angel (1971b) found an average age-at-death for
adult males and females of 34.4 and 29.8 years, respectively, higher than estimates he had
derived for the Upper Paleolithic, thus meeting his expectation of greater “longevity” in
the Neolithic. This increased presence of older, longer-lived adults, he argued, allowed the
society “. . .time to develop and stabilize a culture considerably more complex and richer
than that of the Upper Palaeolithic” (1971b:80). Moreover, for women, the increase in life
expectancy in the Neolithic was important in “. . .allowing more time for childbearing as
well as for training of children” (1971b:80). Thus, his idea of increasing longevity – a
conclusion driven by the presence of more older adults and higher mean age-at-death –
had, in his view, important implications for cultural and social developments in the
earliest farmers of western Asia in comparison with earlier foraging societies from this
region and elsewhere.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
406 Bioarchaeological paleodemography: interpreting age-at-death structures

These early paleodemographic analyses built on interpretations of mean age-at-death and


comparison of relative frequencies of individuals in younger versus older age groups also
provided a tool for documenting both mortality and life expectancy (or longevity), and were
thought to represent the “simplest measure of success biologically speaking” (Angel,
1971b:80). Yet, for many researchers, these simple averages and tabulations were insufficient
for characterizing population structure accurately and did not engage with modern demo-
graphic method and theory. In the view of many authorities (see following), missing from
what bioarchaeologists were doing in their demographic research was formal analysis. In
order to alleviate this shortfall, bioarchaeologists turned to life table analysis, a routine
undertaking by social scientists and others in documenting demographic structures in living
populations.

10.3 Paleodemographers adopt the life table for age structure analysis
Age structure analysis built on the use of life tables was developed by demographers and
actuarial scientists in order to calculate probability of survival to any age, typically starting at
birth (Table 10.1). Because life tables provide the most complete characterization of mortality
in living populations, it made considerable sense for bioarchaeologists deriving their age data
from archaeological skeletal series to apply life table analysis. Bioarchaeologists do this by

Table 10.1 Life table for Northern Ache females. The table is calculated from
survivorship data in Hill & Hurtado (1995) (Adapted from Chamberlain, 2006;
Table 2.2)
x lx dx qx Lx Tx ex

0 1.00 0.27 0.27 4.34 37.35 37.35


5 0.73 0.09 0.12 3.45 33.01 44.97
10 0.64 0.04 0.05 3.13 29.57 45.91
15 0.61 0.02 0.03 2.99 26.43 43.40
20 0.59 0.03 0.06 2.86 23.44 39.86
25 0.56 0.01 0.02 2.75 20.58 37.09
30 0.54 0.04 0.07 2.62 17.84 32.78
35 0.50 0.01 0.01 2.5 15.22 30.19
40 0.50 0.03 0.07 2.40 12.71 25.58
45 0.46 0.04 0.10 2.21 10.31 22.27
50 0.42 0.02 0.05 2.04 8.11 19.35
55 0.40 0.05 0.13 1.86 6.07 15.24
60 0.35 0.07 0.19 1.57 4.20 12.11
65 0.28 0.03 0.11 1.32 2.60 9.41
70 0.25 0.11 0.45 0.97 1.31 5.27
75 0.14 0.14 1.00 0.35 0.35 2.50
80 0.00 0.00 – 0.00 0.00 0.00
x, age at start of five-year interval; lx, survivorship; dx, proportion of deaths; qx,
probablility of death; Lx, average years per person lived within the age interval; Tx, sum of
average years lived within current and remaining age intervals; ex, average years of life
remaining (average life expectancy)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
10.3 Paleodemographers adopt the life table for age structure analysis 407

assuming the age-at-death distribution in the cemetery is equivalent to the dx values in the
life table. Indeed, life table analyses of archaeological skeletal series were rapidly adopted in
the late 1960s and 1970s (Bennett, 1973; Blakely, 1977; Buikstra, 1976a; Kobayashi, 1967;
Lovejoy et al., 1977; Owsley & Bass, 1979; Owsley et al., 1977;Ubelaker, 1974). Acsádi and
Nemeskéri’s (1970) monumental work presented their comprehensive series of life tables,
ranging in time from the late Paleolithic to Medieval Hungary. Importantly, their monograph
was among the first to discuss, in great detail, age and sex estimation methods, quality and
composition of the archaeological skeletal series, and other issues relevant to the study of
ancient populations from skeletal remains.
In the application to North American skeletal series, several key studies revealed the
potential of life table analysis, especially with regard to comparison of multiple populations
within geographical regions. In his classic study, Ubelaker (1974) compared two late prehis-
toric ossuary skeletal series from the Nanjemoy site in the Middle Atlantic region of North
America with a Huron series from Canada (Anderson, 1964). His life table analysis of the
Nanjemoy series revealed relatively few infants and numerous mature adults, suggesting
lower infant mortality and greater adult longevity than in the Huron series. The study was
among the first to apply new methods in age estimation based on histological structures (see
Chapter 2), providing new insights into bone remodeling and age estimation techniques
central to modern paleodemographic analysis.
In lieu of comparisons of demographic profiles derived from archaeological skeletal series,
Weiss (1972, 1973) advocated comparison of archaeological series with model United Nations
life tables (United Nations, 2011). Model United Nations life tables are based on numerous
censuses from around the world and provide mortality data from a wide range of economic,
political, and social circumstances, including those that approximate small scale societies that
are more typically studied by anthropologists (see Weiss’ [1973] model life tables developed
from anthropological field studies). Along this line, Buikstra (1976a) compared estimates of
mortality and survivorship derived from her life table analysis of skeletal remains from the
Middle Woodland Gibson-Klunk mound series (c. 50 BC–AD 400) in the lower Illinois River
valley to model United Nations tables. The general shape of the mortality curve for Gibson-
Klunk suggested a pattern of mortality that approximated recent, modern populations.
What is especially striking about the life table analysis of the Nanjemoy, Gibson-Klunk, and
many other archaeological samples is the paucity of the very young (<5 years) and the very
old (>50 years). This pattern largely reflects poor preservation of the former and inaccuracies
in age estimation methods in the latter (Milner et al., 2008; Walker et al., 1988; and
following). With respect to life table analysis, the apparent underrepresentation of young
juveniles results in considerably skewed (and inaccurate) calculations of the proportion of
deaths (“mortality”) occurring in successive age intervals (dx) as well as the probability of
survival (“survivorship”) of an individual from a cohort that is alive at the beginning of the
age category (lx) (Chamberlain, 2006; Moore et al., 1975). Even skeletal series that were
remarkably large and contained numerous very young juveniles did not seem to represent
realistic population profiles. For example, Howell’s (1982) commentary on Lovejoy and
coworker’s (1977) analysis of the Libben series indicated that “survivorship” of middle- and
older-aged adults as displayed in the Libben life table would have been unrealistically low,
certainly too low to maintain the necessary work force and social context in general for
supporting the community with its dietary and economic needs (Howell, 1982).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
408 Bioarchaeological paleodemography: interpreting age-at-death structures

Despite the newfound use of life tables, a range of problems continued to surface. Angel
(1969) declined analyzing and interpreting population structure based on life table analysis,
largely owing to the vagaries of representation – especially younger juveniles – in archaeo-
logical skeletal assemblages and the inability to account for population growth. The former
has profound influence on mortality calculations, and the latter runs counter to the primary
underlying assumptions of life table analysis, namely population stability. In his mind, the
assumptions about collections of human remains from archaeological contexts did not meet
the key tenets of life table analysis (Frankenberg & Konigsberg, 2006).

10.4 Addressing the assumptions of paleodemography


The following assumptions limit life table analysis as applied to archaeological skeletal series.
First, archaeological life table analysis is built on the assumption that the population from
which the ages are drawn is stable (birth and death rates remain the same over time) and
stationary (the population neither grows nor declines). This assumption is unlikely, as many
settings show either increase or decrease, such as the respective increase in population size in
early farmers and later declines in colonial settings in the Americas (see later). In some
historic-era contexts, written records can provide some perspective on population stability
and change. In Hooton’s (1930) Pecos Pueblo study, historic records provided him with size
estimates, largely showing population decline. Hooton acknowledged that the record of
population change for prehistoric skeletal samples is difficult to track, owing to lack of
comparable records. Although few in number, some methods have been developed for life
table analysis that account for population growth and decline (Bennett, 1973; Frankenberg &
Konigsberg, 2006; Moore et al., 1975; Weiss, 1973). However, even these approaches have
limitations.
Compounding the unresolvable question of population stability is the fact that the distri-
bution of age-at-death in an archaeological series of remains is biased by the variation in
probability of death across different age classes (Jackes, 2011). In this regard, death rates are
not uniform across age and sex, influencing change in a population’s age–sex structure
through time. For example, perinatal death rates for males are higher than for females, and
the shift from consumption of breast milk to adult diet causes a spike in infant mortality (see
Chapter 3). Simply, different age groups are subject to different circumstances of frailty, and
hence, lack uniformity in patterns of death.
Second, life table analysis assumes that the collection of tabled skeletons represents the true
biological population from which the collection is derived, and that no part of the population
is underrepresented from either poor preservation (especially young juveniles) or from differ-
ential mortuary treatment of the deceased (e.g., separating juveniles from adults). The influ-
ence of representation in any skeletal series from an archaeological context is now well
known, especially deriving from differential preservation. A range of archaeological contexts
show poor preservation of juvenile remains and differential preservation generally, often in
relation to soil chemistry (Bello et al., 2006; Buikstra, 1976a; Gordon & Buikstra, 1981; Jackes,
2011; Meindl & Russell, 1998; Pinhasi & Bourbou, 2008; Walker, 1995; Walker et al., 1988;
Wittwer-Backofen et al., 2008). These circumstances can influence a researcher’s ability to
estimate age and sex, especially when fragile skeletal elements such as the pubic bone are not
preserved.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
10.4 Addressing the assumptions of paleodemography 409

The choice of archaeological recovery methods and site sampling contributes to variable
sample representation. The excavation at Neolithic Çatalhöyük reveals the importance of
the presence of a clearly defined excavation protocol in determining what is or is not
recovered. Angel (1971b), for example, reported a very low presence of young juvenile
remains excavated by Mellaart. Later excavations produced an unusually high number of
juvenile remains, well over half of the collection, including numerous neonates (Hillson et al.,
2013). The remains examined by Angel were recovered using an excavation protocol involv-
ing rapid shoveling of considerable amounts of fill, whereas the more recent excavations
involved highly detailed recovery techniques, such as screening. Thus, the demographic
analysis based on the latter will likely be more accurate and representative than that based
on the former.
Juvenile remains can also be underrepresented under circumstances where the mortuary
treatment is different from that of adults. For example, shallower burial of juveniles than
adults in some instances in Hungary resulted in poorer preservation of the former relative to
the latter (Acsádi & Nemeskéri, 1970; and see Bello et al., 2006). Moreover, entire age groups –
such as young juveniles – may be missing from cemeteries because they are interred separ-
ately from the main population (Pinhasi & Bourbou, 2008; Schwartz et al., 2010, 2012).
In addition, the completeness of excavation of a mortuary context can play a critical role
in which segments of a population are represented in a study series. Buikstra (1976a) notes
that late nineteenth-century excavations at the Pete Klunk mound group in the lower Illinois
River valley produced only 16 burials, with an average of three skeletons per mound. In
contrast, later excavations involving complete recovery and excavation produced an average
of 38 burials per mound.
Third, life table analysis assumes that the ages of the individuals comprising the
collection of remains are accurate. Indeed, it is this assumption that is so very crucial
to the meaning of ancient age structures. The clear breaking point in discussions of
paleodemography is linked to the controversial article by Bocquet-Appel and Masset
(1982) who concluded from their analysis of the record that, “It is not the information
provided by the age indicators which is dangerous; it is the fruitless need to make them
say more than they can” (1982:332). While concluding that ancient skeletons are
certainly a fund of information for understanding past population biology, they single
out the record of demography as a dead end, bidding “farewell to paleodemography”
(1982:332). Simply, age estimation methods as applied to archaeological skeletons are
too imprecise and fraught with problems to provide anything meaningful about age
structures except in the very crudest way. Moreover, Bocquet-Appel and Masset (1982)
concluded that age structures provided by paleodemographers simply represent artifacts
of the age-at-death composition of the original reference sample used to estimate age-
at-death in the target series. For example, if the age estimations for an archaeological
series were made using the Todd (1920) pubic symphysis method, then the age-at-death
profile simply approximates the series Todd used in the development of his method.
Thus, in the view of Bocquet-Appel and Masset, the average age-at-death of a skeletal
series represents a combination of real biological processes of aging and the reference
sample age structure. If this critique held true, then paleodemography has fundamental
(and fatal) flaws.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
410 Bioarchaeological paleodemography: interpreting age-at-death structures

10.5 New solutions to interpreting age-at-death profiles in archaeological


skeletal series: it is really mostly about fertility not mortality
In the short run, bioarchaeologists were shocked by Bocquet-Appel and Masset’s “farewell
to paleodemography” (1982). Some authorities expressed indignation and extreme doubt
that age estimations and mortality profiles are influenced by the age structure of the
reference sample. After all, do not all humans age and/or senesce at the same rate and in
the same manner? Van Gerven and Armelagos (1983) reassessed Bocquet-Appel and Masset’s
assertion that the Nubian series was biased by the Korean War dead sample that McKern and
Stewart (1957) had used to develop their age estimation method. In order to address the
conclusion drawn by Bocquet-Appel and Masset, Van Gerven and Armelagos compared age-
at-death of well-preserved Nubian skeletal remains from Wadi Halfa (n¼201) and Kulubnarti
(n¼162) using Todd’s (1920) and Brooks’s (1955) methods of age estimation for the pubic
symphysis. Using the Kolmogorov–Smirnov two-sample test for cumulative frequencies and
applied to cumulative mortality curves, they documented a strong, highly significant difference
between the three series, suggesting that the age profiles of archaeological series are not influ-
enced by the reference sample. Based on their analysis, the study presents findings to suggest
that at least part of the record of paleodemographic research warrants further attention.
Similarly, others made the case that Bocquet-Appel and Masset’s (1982) criticisms were extreme
and premature. That is, use of multiple age indicators, selective use of age estimation methods
with known accuracy, use of population-specific age estimation methods, development of
methods for aging older individuals, and standardization of age estimation methods, a direc-
tion for research supporting the continued importance of paleodemography would be provided
(Buikstra & Konigsberg, 1985; Lovejoy et al., 1985). Arguably, while the death of paleodemo-
graphy was announced prematurely by Bocquet-Appel and Masset, fundamental elements of
age structure analysis were in need of attention (Hoppa & Vaupel, 2002; Konigsberg &
Frankenberg, 1992, 1994, 2002).
Regardless of the ton of comments, either for or against abandoning paleodemography
altogether, Bocquet-Appel and Masset (1982) set into motion a debate about the relative value
of paleodemography and the need for new methods for reconstructing and interpreting
ancient age structures in past societies. Their work served as a call for developing methods
that would address shortcomings in age estimation methods as well as elucidating how death
assemblages influence reconstructions of demographic composition. It made clear that age
profiles derived from censuses of living populations and death assemblages are simply not the
same, with the latter subject to biases and circumstances not present in the living. Censuses of
living populations are based on known ages, whereas censuses of dead individuals from a
collection are based on estimates fraught with problems, owing to use of methods, which for
the most part, assume uniformity across age groups and across human populations in aging
and skeletal indicators of age. Moreover, it is simply untenable to continue attempting to fit
age structures of archaeological skeletal series into patterns that may or may not be predicted
by model life tables developed from entirely different environmental, social, and economic
circumstances. Basically, while the life table has analytical merits for predicting length
of life in living populations, it has limited usefulness in paleodemographic investigation
(Hoppa, 2002).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
10.5 New solutions to interpreting age-at-death profiles in archaeological skeletal series 411

Not surprisingly, almost immediately following Bocquet-Appel and Masset’s (1982) publi-
cation, bioarchaeology saw the slowdown of the publication of paleodemographic reports and
life tables, reflecting the growing recognition that considerably more thought had to be put
into identifying the strengths and weaknesses of deriving mortality profiles from archaeo-
logical skeletal series. As is true of any maturing science, Bocquet-Appel and Masset’s
negative assessment had a positive, constructive outcome. Ultimately, it caused researchers
to undertake thoughtful considerations of what age-at-death and age structure reveal about
population dynamics when these population dynamics rely on collections of dead people
versus populations of living individuals.
Tackling this problem directly, Sattenspiel and Harpending (1983) addressed the question:
Is it appropriate to use mean age-at-death in a skeletal series from a single site or collection of
sites as an appropriate estimate of life expectancy at birth? Put another way, does hypothet-
ical population x with a mean age-at-death of 19.6 have a lower life expectancy at birth than
hypothetical population y with a mean age-at-death of 22.3 years? A reading of much of the
bioarchaeological paleodemographic literature would certainly seem to affirm this assump-
tion, and life table analysis appeared to support the notion as well. Sattenspiel and Harpend-
ing (1983) reminded anthropologists that the procedure is a valid one only if populations are
stationary – that birth and death rates remain stable and population size remains unchanged,
neither increasing nor decreasing. They made the compelling case that this assumption is
rarely met in living populations. Thus, by inference, could that be the case for past popula-
tions? Does mean age-at-death reflect life expectancy in past populations? The answer is no.
Using simple algebraic analysis, they analyzed a hypothetical stable skeletal population
having a uniform age-specific death rate. Specifically, the analysis showed that even a small
change in the birth rate results in a proportional change in mean age-at-death, similar to what
Coale (1972) found in his examination of the influence of fertility and mortality changes on
age distribution in living populations. Mean age-at-death in a series of skeletons is strongly
influenced by fertility and not mortality – mean age-at-death is the inverse of birth rate.
Therefore, if mean age-at-death is declining, then birth rate is likely increasing. This break-
through in paleodemography, therefore, shows that fertility has a considerably stronger
influence on the distribution of deaths in a skeletal series than does mortality. Certainly, this
notion of a death assemblage being more informative about fertility than mortality is counter-
intuitive, so much so that it is now far more common in the literature to see bioarchaeologists
interpreting skeletal samples as records of mortality (Coppa et al., 1995; Nagaoka & Hirata,
2007; Nagaoka et al., 2006; Sullivan, 2004).
When one considers that fertility-focused paleodemography is really more biological than
demographic, the concept is far more intuitive. One need only consider the classic population
pyramid with the oldest at the peak and the youngest at the base (Figure 10.2). In virtually all
mammals, the rate of dying is highest at the pyramid’s base and peak, with the youngest and
the oldest dying in higher proportions than the other age groups. However, fertility happens
only at the base of the pyramid, while mortality is spread throughout the rest of the pyramid.
This means that when there is high fertility, the base of the pyramid is wide and the rest of the
pyramid becomes progressively narrower. When there is low fertility, the base is narrow and
the ages are distributed higher on the pyramid. High fertility, therefore, makes for a young
population. Thus, no matter how counterintuitive fertility-based paleodemography may be,
death assemblages tell far more about fertility and birth rate than mortality and death rate.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
412 Bioarchaeological paleodemography: interpreting age-at-death structures

(a) Melanesia

75+
70−74
65−69
60−64
55−59
50−54
Age groups 45−49
40−44
35−39 Male
30−34
Female
25−29
20−24
15−19
10−14
5−9
0−4
10 5 0 5 10
Percent population

(b) Polynesia

75+
70−74
65−69
60−64
55−59
50−54
Age groups

45−49
40−44
35−39 Male
30−34
Female
25−29
20−24
15−19
10−14
5−9
0−4
10 5 0 5 10
Percent population

(c) Australia

75+
70−74
65−69
60−64
55−59
50−54
Age groups

45−49
40−44
35−39 Male
30−34
Female
25−29
20−24
15−19
10−14
5−9
0−4
10 5 0 5 10
Percent population

Figure 10.2 Population pyramids for Melanesia (a), Polynesia (b), and Australia (c) in 2010.
Note the demographic discrepancies in age groups between populations. Higher fertility
populations have a relatively wide base. (Graph produced from data published by United
Nations, Department of Economic and Social Affairs, Population Division, 2012.)

Simply, as noted by Robert McCaa, “Common sense tells us that the age structure of deaths is
completely determined by mortality, but common sense is wrong” (2002:98).
Sattenspiel and Harpending (1983) examined several published studies and derived new
interpretations that make considerably more sense than what had been reported by Acsádi

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
10.5 New solutions to interpreting age-at-death profiles in archaeological skeletal series 413

and Nemeskéri (1970) for skeletal series from Maghreb (Paleolithic), Alsónémedi (Copper Age),
Intercisa and Brigetio (Roman period), Sopronköhida (early Medieval), and Hungary (late
Medieval), by Green and coworkers (1974) for Nubia, and by Angel (1971a) for Greece. Their
birth rate revealed a temporal pattern indicating population dynamics and growth (or decline)
that was very different from the conclusions drawn by earlier investigators. Similarly,
reanalysis of the Dickson Mounds age structure reveals the likelihood that the decline in
mean age-at-death is best interpreted as an increase in birth rate and fertility, rather than
reflecting a decline in life expectancy and increased mortality (Johansson & Horowitz, 1986).
The result of the Sattenspiel and Harpending (1983) study was to make the strong case that
mean age-at-death is a simple inverse of birthrate: the lower the mean age-at-death, the
greater the birthrate (and see Acsádi & Nemeskéri, 1970; Bocquet-Appel & Masset, 1982;
Johansson & Horowitz, 1986; McCaa, 2002; Milner et al., 1989; Wood et al., 1992). However,
this fundamental development in bioarchaeological demography still does not handle the
problem facing most bioarchaeologists in assessing growth (or decline) of a population – the
near-universal underrepresentation of young juveniles in archaeological death assemblages.
The link between age structure and birth rate is clear, but how to accommodate the very
thorny issue of incompleteness of skeletal samples, especially regarding the frequent under-
representation of young juveniles, is less straightforward. This is no small matter if one
expects to apply fertility models to explain key population events, such as (and perhaps
among the most important) the dramatic and global rise in population beginning in the
Neolithic with the shift from foraging to farming. The archaeological community has argued
the shift from foraging to farming was accompanied by an increase in population size owing
to increased growth rates, mostly due to the rise in fertility. Moreover, ethnographic analogies
indicate that horticulturalists and incipient farmers have higher fertility and birthrates than
foragers (Bentley et al., 1993; Gage & DeWitte, 2009; Hewlett, 1991; Leslie et al., 1999). How,
then, do we tackle the problem of missing juveniles and still estimate patterns of fertility and
birthrate?
Acknowledging the issue of underrepresentation and still being able to explain change in
population structure, especially with regard to understanding the dynamics of population
increase in recent human evolution, Buikstra and coworkers (1986) offered a case study from
the Illinois River valley, where population increase was clearly a phenomenon of later
prehistory and was almost certainly linked to the introduction of and increased dependence
on maize agriculture. Following Bocquet-Appel’s lead (Bocquet-Appel & Masset, 1982), in
order to circumvent the clear underrepresentation of young juveniles, they used a fertility
estimator – a juvenility index calculated by the proportion of deaths equal to and exceeding the
age of 30 years (D30þ) to deaths equal to and exceeding the age of five years (D5þ). Consistent
with expectations articulated by Sattenspiel & Harpending (1983), there is a decrease in the
D30þ/D5þ ratio from the Middle Woodland through the Late Mississippian periods, reflecting
an increasingly younger population, a factor driven largely by higher birthrates and increased
fertility overall in later populations compared with earlier populations (Figure 10.3). There-
after, a considerable record of use of the juvenility index, either the one proposed by Buikstra
and coworkers or modifications thereof, reveals a consistent pattern in virtually every setting
examined, especially following upon the expected outcomes associated with either the
adoption or intensification of agriculture. In specific regional settings, this general pattern
of elevated fertility and birthrate has now been well documented in the Near East (Guerrero

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
414 Bioarchaeological paleodemography: interpreting age-at-death structures

0.90
0.85
0.80
0.75
0.70
D30+/D5+

0.65
0.60
0.55
0.50
0.45
0.40
0.35

LW
LW

LW

s
LW

LW
W

is
M

M
o.

ild
n

er

s
o.

er
to

ld
C

st
un

h
C

el

dd

hi
Sc
Ko
ke
Kl
ke

Sc
Le
Pi
n/
Pi

so
ib
G

Skeletal Series
Figure 10.3 Juvenility index or the proportion of deaths, D30þ/D5þ, in prehistoric
west-central Illinois skeletons. Middle Woodland sites: Pike County MW and Gibson/Klunk
MW; Late Woodland sites: Pike County LW, Helton LW, Koster LW, Ledders LW, Schild
LW; and Mississippian sites: Schild Knolls Miss. (Adapted from Buikstra et al., 1986;
reproduced with permission of author and the Society for American Antiquity.)

et al., 2008; Hershkovitz & Gopher, 2008; Littleton, 2011), Europe (Bocquet-Appel, 2002;
Bocquet-Appel & Naji, 2006; Jackes & Meiklejohn, 2008) (Figure 10.4), South and Southeast
Asia (Domett & Oxenham, 2010; Pietrusewsky & Douglas, 2002a, 2002b; Robbins Schug,
2011a, 2011b), various settings in North America (Buikstra et al., 1986; Kohler et al., 2008;
Larsen et al., 2007), and North America generally (Bocquet-Appel & Naji, 2006; Bocquet-
Appel et al., 2008) (Figure 10.5). This remarkably consistent pattern, showing an abrupt
increase in the proportion of juveniles in archaeological death assemblages, provides an
essential element for understanding one of the most important adaptive transitions in human
evolution, namely the foraging-to-farming transition, or what has also been called the
Neolithic demographic transition (Bocquet-Appel, 2002, 2008, 2011) (Figure 10.6). This abrupt
spike cannot be related simply to differences in preservation of skeletons or mortuary
behavior. Rather, it signals a worldwide and rapid transition to agriculture, at least in those
settings studied. Importantly, this development in modern bioarchaeological paleodemogra-
phy adds important context for understanding the various indicators of morbidity, dietary,
and lifestyle issues discussed elsewhere in this book and indeed the entire package of the
foraging-to-farming transition.
All of the aforementioned research speaks to the development and growth of food produc-
tion and the expected link to increased fertility. Likely, the underlying theme is the increased
availability of stored foods made possible by storage of grains and the shift to a practice of
producing weaning foods from cooked grains and consumption of soft mushes by young
juveniles. However, it is important to point out that the shift to agricultural production and
intensification is not always linked to an increase in birthrate or population growth more

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
10.5 New solutions to interpreting age-at-death profiles in archaeological skeletal series 415

0.7

La_
0.6 Cal
Hei

Bad
0.5 Par
Nor
Ded
Ajd
Haz
Cau
0.4 Len Viln Sch
D19+/D5+

Les
Ais Mor
Nie
Bru Der
Bre Wan Gro
Pon Fon Cas Ca Pie
0.3 Bau Cer
Ved∑ Son Cov Bel
Col∑ Cha Villan Gou
Yas Mont
Mal Jun
Stu Rea Vik
Ta∑ Nit Bro Loi Eyb
0.2 Sam Mon Pec
Rut
Die∑ Cab Lar
Moi∑ ∑
Ska Ait Tre Vill
Vedr Oct Ave Cen
Mai
0.1
zen Dic

0.0
0 00

00

00

00
0

0
00

00

00

00

10

20

30

40
–4

–3

–2

–1

dt (years)
Figure 10.4 Proportion of deaths, D19þ/D5þ, among 68 prehistoric European cemeteries,
plotted to chronological distance from the Neolithic diffusion front (dt, the time elapsed
from the beginning of farming at each locality). The dashed line represents the expected
proportion of subadults if the growth rate is zero. Subadult prevalence data indicate an
increase in growth rate associated with the Neolithic transition. (From Bocquet-Appel,
2002; reproduced with permission of author and University of Chicago Press.)

generally. There certainly is a link between farming and population increase for most settings
studied by bioarchaeologists. Nevertheless, in some settings where agricultural intensification
has taken place, there have been reductions in fertility owing to generally deteriorating
circumstances and life conditions. The increase in morbidity associated with these deterior-
ating circumstances is well outlined elsewhere (see Chapter 3). Suffice it to say that increased
prevalence of infection and infectious disease owing to living in closer, more crowded
conditions of semi-permanent or permanent settlements, decreased nutritional quality, and
other factors must have contributed to increased mortality, thus partially or entirely offsetting
the elevated fertility rates now documented in the bioarchaeological record. This underscores
the important point that population dynamics as they relate to size is a balance between
fertility and mortality.
Colonial Florida is one such setting where agriculture intensified, but depressed living
circumstances and elevated mortality contributed to a decline and eventual extinction of a
population (Larsen, Griffin et al., 2001; Larsen et al., 2007; Russell et al., 1990). In this setting,
key dietary and behavioral changes contributed to an increase in infectious disease, increased
focus on maize, loss of marine resources, and extraordinary labor demands placed upon

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
416 Bioarchaeological paleodemography: interpreting age-at-death structures

0.6

0.5

0.4
D19+/D5+

0.3

0.2

0.1
0 00 00 00 0 0 0 00 00 00
00 0 0 0 00 00 10 20 30
-6 -5 -4 -3 -2 -1
dt (years)
Figure 10.5 Proportion of deaths, D15þ/D5þ, among 62 North American samples, relative
to chronological distance from the Neolithic diffusion front (dt). The dashed line represents
the expected proportion of subadults if the growth rate is zero. (From Bocquet-Appel & Naji,
2006; reproduced with permission of authors and University of Chicago Press.)

0.5

0.4

0.3
15p5

0.2

0.1

0.0
0 0 0 00 0 00 00 00 00
00 00 00 0 10 20 30 40
-4 -3 -2 -1
dt (years)
Figure 10.6 Proportion of deaths, D15þ/D5þ, among 133 Northern Hemisphere cemeteries,
relative to chronological distance from the Neolithic diffusion front (dt). A significant
increase in population is observed during the Neolithic period. The dashed line represents
the expected proportion of subadults if the growth rate is zero. (From Bocquet-Appel, 2011;
reproduced with permission from American Association for the Advancement of Science &
from Springer Science and Business Media.)

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
10.5 New solutions to interpreting age-at-death profiles in archaeological skeletal series 417

0.8

0.7

0.6

0.5
D30+ / D5+

0.4

0.3

0.2

0.1

0
PP PA EM LM
Guale Skeletal Series
Figure 10.7 Proportion of deaths, D30þ/D5þ, among Prehistoric preagricultural (PP),
Prehistoric agricultural (PA), Early Mission (EM), and Late Mission (LM) Guale populations,
Spanish Florida. These juvenility indices (D30þ/D5þ) reveal a significant increase in the
proportion of older adults to subadults from Prehistoric and Early Mission to Late Mission
periods, which is indicative of reduced population fertility.

native societies by European powers. During the later seventeenth century, combined pres-
sures from slaving forays into Spanish Florida by native enemies and encroachment south-
ward by British interests led to considerable declines in the population. We know, therefore,
that increased mortality, morbidity, and decaying conditions likely contributed to increased
deaths and declining population size. Occurring over a period of decades, this would have
depressed the size of the population, which, indeed, has been well documented in historic
sources (Worth, 2001, 2007). However, every demographer knows that mortality represents
only one side of the demographic coin – fertility rates also play an important role. In order to
test the hypothesis that fertility and birthrates declined in native populations of Spanish
Florida, application of the juvenility index (D30þ/D5þ) was calculated for a skeletal series
representing the Guale tribal group, including prehistoric foragers, prehistoric farmers, early
mission, and late mission populations. The ratios for the prehistoric foragers, prehistoric
farmers, and early mission Native Americans are statistically indistinguishable, but show
some fluctuation in the juvenility index (0.3790, 0.3500, 0.2823, respectively; Figure 10.7). In
particular, there is a suggestion of increased fertility in the late prehistoric period when maize
farming was present, followed by a further increase in fertility in the earliest mission
occupation. However, in the late mission sample, there are very few children and a consider-
able number of older adults, resulting in a juvenility index of 0.7623 – a significant depression
in fertility. It is also during this period that there is a highly elevated morbidity profile (e.g.,
dental caries, cribra orbitalia, porotic hyperostosis) (Larsen, Griffin et al., 2001; Larsen et al.,
2007). Moreover, historic records on population size, living conditions, increased population

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
418 Bioarchaeological paleodemography: interpreting age-at-death structures

nucleation, and peak disruption are consistent with the notion that fertility declined. Coupled
with increased mortality and morbidity, it is little wonder that the population in Spanish-
controlled Florida had reduced from hundreds reported to just scores of individuals, all within
a period of a few decades. This outcome of declining fertility per the juvenility index is similar
to the pattern identified in early contact-era Huron ossuary samples (Kleinburg, Uxbridge,
Ossossané) from early seventeenth-century Ontario (Jackes, 1994) and the colonial north
coast of Peru (Klaus & Tam, 2009). While the pattern of colonialism is different in the Spanish
Florida, Spanish Peru, and Ontario settings, the outcome of contact and population collapse is
clearly revealed in estimates showing declining fertility.

10.6 The elephant in the room: age estimates in archaeological skeletons


Bocquet-Appel and Masset’s (1982) highly critical overview of paleodemography included a
scathing attack on the status quo of age estimation methods as they are applied to archaeo-
logical skeletons, especially the standard methods for the pubic symphysis developed by
physical anthropologists for individuals of known age-at-death (McKern & Stewart, 1957;
Todd, 1920). They argued that the ages derived for the reference skeletons (Todd, 1920)
strongly influence the age structure of the target skeletons (archaeological series). Indeed,
later studies have shown the strong effect of the reference sample on age estimates and
demographic profiles (Hoppa & Vaupel, 2002; Jackes, 2011; Konigsberg & Frankenberg, 1992,
1994; Milner et al., 2008). However, immediately following the Bocquet-Appel and Masset
(1982) critique, the response from the bioarchaeological community ranged from dismissive to
disgust. After all, authorities asked, how could the methods focusing on age changes in the
pubic symphysis by such careful workers and luminaries in the discipline as Thomas McKern,
T. Dale Stewart, and T. Wingate Todd and commonly applied to archaeological remains be
challenged in such a negative manner?
Despite the resistance to the notion that ages of archaeological skeletons partly reflect the
age structure of the reference populations used to develop those methods, the critique sparked
a debate, leading a number of individuals to reassess the issue. There is now a consensus that
estimates of age in archaeological series are influenced by the reference skeletal sample
(Hoppa & Vaupel, 2002). Thus, it is essential that age estimation methods are more reliable
and better validated. One solution has been to use multiple age indicators (Lovejoy et al.,
1985). Nonetheless, the use of multiple indicators does not provide a solution necessarily, as at
least some of the indicators are derived from reference samples. Dental seriation based on
occlusal surface wear in juveniles is not based on a reference sample (Miles, 1963), but even
with this method, the increasingly erratic wear occurring after the age of 40 years limits the
ability to estimate age-at-death for often significant portions of archaeological skeletal series
(Gilmore & Grote, 2012). Similarly, cranial suture closure provides some insight into age in the
first three decades of adulthood, but is uninformative in later years (Lynnerup & Jacobsen,
2003; Milner & Boldsen, 2012a, 2012b, 2012c; but see Wittwer-Backofen et al., 2008).
Moreover, with few exceptions age estimation methods tend to be within a narrow range
for juveniles and young adults, and considerably broader ranges for adults generally, espe-
cially underestimating age-at-death for individuals more than 50 years old (Milner & Boldsen,
2012a, 2012b; 2012c; Milner et al., 2008; Nawrocki, 2010; Wittwer-Backofen et al. 2008).

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
10.7 Summary and conclusions 419

Pubic Symphysis
Actual Dashes and Estimated (Lines) Ages
100
80
60
40
20

0 50 100 150 200 250


Individuals from Youngest to Oldest

Figure 10.8 Age range estimates for modern day skeletons (vertical lines) based on pubic
symphysis standards. The black dashes represent the actual age-at-death of the individual.
(From Milner & Boldsen, 2012c; reproduced with permission of authors and John Wiley &
Sons, Inc.)

A newly emerging method advocates using the large suite of skeletal characteristics that
undergo progressive age changes, and establishing ages of transition over the course of these
changes for each trait, such as for the pubic symphysis (Boldsen et al., 2002; Godde & Hens,
2012; Milner & Boldsen, 2012a, 2012b, 2012c). The distinctions in age for individual skeletal
traits provide minimal information on age (e.g., “young” versus “old”), but as a collective
approach, transition analysis provides excellent results, especially when numerous traits are
employed simultaneously to assess the transitions (Milner & Boldsen, 2012c; Weise et al.,
2009). Validation of the method suggests that this approach extends the limit of more precise
age estimates from 40 years to more than 60 years old, thus providing a more meaningful
demographic reconstruction of age-at-death profiles in archaeological skeletal series (Milner &
Boldsen, 2012c) (Figure 10.8). Use of this method, combined with probabilistic (Bayesian)
approaches to age distribution of the target population independent of the reference popula-
tion, presents new opportunities for improving the accuracy of age estimates in order to
reconstruct and interpret past demographic patterns in a more informed manner (Boldsen
et al., 2002; Godde & Hens, 2012; Hoppa & Vaupel, 2002; Milner & Boldsen, 2012c).

10.7 Summary and conclusions


The record of death provides important insight into conditions of the living. We now
understand that to talk about mean age-at-death in relation to mortality and life expectancy
is unrealistic. Emerging practice reveals that age structures of archaeological death assem-
blages provide insight into population structure, especially fertility rates, in some key ways.
Moreover, like all areas addressed in this book, results need to be examined in the light of full
social, cultural, economic, and environmental contexts, asking the question: Do the results of
the study make sense in light of the context? This has certainly proven to be the case in
specific settings like Spanish Florida where predictions of declining birthrate and an aging

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
420 Bioarchaeological paleodemography: interpreting age-at-death structures

population are consistent with the record. Similarly, the growth in population in the American
Midwestern Illinois River valley and elevated birthrates identified in this well-studied region
are consistent with expectations of birthrates in early farming populations that exhibit
increasing settlement size and distribution. There is some evidence to suggest specific patterns
of age compositions in the comparison of catastrophic and attritional death assemblages
(Gowland & Chamberlain, 2005; Margerison & Knüsel, 2002; but see Paine, 2000; Roberts and
Grauer, 2001). However, it falls on the shoulders of the investigator to fully understand the
social, economic, political, behavioral, and taphonomic contexts when interpreting age
structures.
The study of historical skeletal and demographic profiles involving catastrophic mortality
events clearly shows the importance of understanding the circumstances surrounding death
and who is or is not at risk of dying (Grayson, 1990, 1993; Williams, 2009). An example of
this can be seen in the infamous disaster of starvation and death in the Donner Party, a group
of pioneers who in 1846 were stranded in the Sierra Nevada Mountains when snow made it
impossible for them to reach their destination in California. The party of men, women, and
children ran out of food, resulting in the deaths of 35 of 83 individuals due to starvation
(Grayson, 1990, 1993). The demographic record provides important perspective on frailty and
those most vulnerable to death. Modern demographic data would predict that male mortality
would be higher than female mortality. Indeed, that was the case in this setting where the
number of starvation deaths was greater for males than females, especially younger males. In
addition, males died earlier than females in the disaster episode. Especially striking about the
age profiles were the apparent risks associated with not having kin – those without kin died
earlier during the course of the disaster than those with kin (Grayson, 1990, 1993). Thus, the
survival in these difficult circumstances was related to sex, age, and kin group size (and see
Grayson, 1996 for a similar catastrophic context in western North America).
Analysis of the demographic profile of victims of catastrophic flooding following the
Johnstown Flood of 1889 provides a completely different demographic profile than the
starvation deaths in western North America (Williams, 2009). In this setting, deaths occurred
mostly within hours following the bursting of a dam located upriver from Johnstown,
Pennsylvania. Flood waters reached the town virtually without warning, resulting in the
deaths of about 2200 men, women, and children, roughly 10% of the town’s inhabitants.
Analysis of demographic profiles derived from archival records revealed important perspec-
tive on vulnerability in relation to both natural and culturally mediated hazards. In particular,
the rapidly occurring deaths in Johnstown were more heavily weighted toward females than
males, especially younger females (Figure 10.9). Williams (2009) suggests that circumstances
relating to geographical, social, or biological vulnerabilities were likely important in influ-
encing the significant differences between men and women as pertaining to death and injury
in this setting. In particular, the higher mortality rates in women were likely related to more
women being at home than men at the time the floodwaters struck the town and their
considerably bulkier clothing, rendering women less able to maneuver in the flooded setting.
Other factors may also explain the sex differences in deaths. That is, an analysis of a variety
of circumstances involving rapid catastrophic deaths in a range of settings globally reveals
that women (and children) have higher death rates, perhaps due to less endurance and
strength. In addition to these sex differences, there was considerably greater mortality in
children than adults, and especially in individuals less than five years of age. Like most

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
10.7 Summary and conclusions 421

35
Victims of the Johnstown Flood
Females
30
Males
All
25
Percentage

20

15

10

0
0−9 10−19 20−29 30−39 40−49 50−59 60−69 70−79 80−89 90+
Age (in years)
Figure 10.9 The age distribution for Johnstown, Pennsylvania flood victims demonstrates
that adult females (20–40 years old) were more vulnerable to the effects of the flood than
their adult male counterparts. (Data from Williams, 2009; reproduced with permission of
author.)

disaster settings, children are more vulnerable owing to their smaller size, and minimal
stamina and strength (Glass et al., 1977; Seaman et al., 1984). These studies of historical
populations underscore the importance of examining the natural, social, cultural, and eco-
nomic contexts in the analysis of demographic profiles. As per Richard Paine’s (2000:182)
advice, “. . .to make robust conclusion from paleodemographic data, we must develop explicit
models of how cultural and biological processes may manifest themselves in the skeletal
record.”
Many caveats surround paleodemographic study. Contrary to earlier assumptions about
manner and rate of aging and senescence, these processes vary both within and across human
populations. Therefore, it is reasonable to state that the ages of reference populations have an
influence on age estimation and thus age structure of target (archaeological) populations.
Moreover, the age-at-death profiles generated – no matter how accurate – are not the same in
composition as those generated for living populations. The developments in what can and
cannot be said about aging and age structures of archaeological skeletal samples and the
impressive advances in theory and method behind their interpretation are cause for welcom-
ing a new and invigorated paleodemography – earlier reports of its death were greatly
exaggerated.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 13 Jan 2017 at 19:53:34, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.012
11 Bioarchaeology: skeletons in context

11.1 Framing the contextual record


Two elements stood out for me as I wrote this book: (1) human remains provide a remarkably
rich record of the human experience, and (2) that record is further enriched in consideration of
the context from which these remains derive. This book emphasizes the record that human
remains provide for characterizing a broad range of issues about the past, but they have
limited value without the development of a comprehensive understanding of their broader
context. There are multiple levels of context, including the manner in which the remains were
recovered via excavation and during curation, the cultural and social processes and events of
the interred and of the surviving members of the community during the burial event, the
environmental setting in which the individual lived, and behavioral reconstruction of the
interred, all of which are informed by ethnographic, historical, archaeological, and other
sources.
A basic issue in bioarchaeology that can be assessed using contextual information is the
extent to which assemblages of archaeological skeletal series represent the once-living
populations from which they were drawn. For example, does the high prevalence of periosteal
reactions, tuberculosis, or osteoarthritis in a series simply reflect a bias toward inclusion of
only older individuals who accumulated more lesions over their long lives, or might the series
reflect those who were the least frail and survived and recovered from biological stress? Are
there individuals missing from the record that limits the kinds of questions we can ask about
it? These are fundamental concerns and questions for virtually any scale of study, whether
involving a single skeleton, a single cemetery, or a collection of cemeteries representing
multiple communities. It is well understood that human remains from archaeological sites
may represent a biased representation of a once-living group from any single time. After all,
sets of human remains from an archaeological locality are almost never snapshots of bio-
logical populations. Rather, they are largely cumulative aggregates of remains, often contain-
ing multiple generations of individuals spanning hundreds or thousands of years (Cadien
et al., 1974). These aggregates typically include individuals who died at different times and
under very different circumstances, including, for example, differing conditions of illness and
traumatic death. One has only to think of one’s local town or city cemetery and peruse the
gravestones of the deceased to speculate on the remarkable range of circumstances that
underlined their deaths. Archaeological death assemblages comprise individuals with varying
life histories, some of which are punctuated by illnesses and events that predisposed them to
elevated frailty and earlier death.
Because of the potential for biological selectivity, individuals included in archaeological
skeletal series may have depressed growth velocity, elevated mortality, and elevated morbidity

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 13:55:51, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.013
11.1 Framing the contextual record 423

and sickness in comparison with the larger population from which they are drawn. Moreover,
this does not even take into account the varied cultural and social circumstances that
determined why a person’s remains were interred in the cemetery, be it their age, gender, or
level of wealth or status in complex or otherwise ranked societies. Death assemblages,
therefore, represent a complex composite of individual circumstances that require careful
consideration.
Wood and coworkers evaluated some potential issues of biological selectivity in their
landmark 1992 study (Wood et al., 1992). They argued that documentation and interpretation
of health levels derived from archaeological human remains is neither a straightforward
exercise nor an intuitive process. That is, temporal increases in skeletal lesion prevalence
(e.g., periosteal reactions, enamel defects) in a series may not be a simple measure of declining
community health. The opposite may be the case instead: increasing numbers of infectious
lesions may indicate an improvement in health. Wood and coworkers contend that “the
frequency of active lesions in a skeletal sample is greater than the fraction of affected
individuals in the living population from which the sample was drawn” (1992:349), as
individuals suffering from such afflictions would be at an increased risk of death compared
to the population as a whole. They also contend that individuals within a population having
lesions may be drawn from the biologically advantaged group rather than the disadvantaged
group, as they were able to survive one or more episodes of chronic stress (e.g., disease,
malnutrition). On the one hand, the implications of this “osteological paradox” are important
in that skeletal evidence for declining health in populations undergoing transitions – such as
from foraging to farming – might just as easily be interpreted as representing improvement in
health. On the other hand, Wood and collaborators (1992) did not intend for future gener-
ations of bioarchaeologists to uncritically throw their hands up in the air, and cry “it’s an
osteological paradox – more pathology, better health!” every time they encountered a
population with a high prevalence of lesions. Quite the opposite: the intention is to highlight
the point that bias, selectivity, and scrutiny of the contextual factors underlying an archaeo-
logical skeletal series must be taken into account. Contemporary bioarchaeology has greatly
benefited from their caution.
Therefore, how might such issues of selectivity and bias be best addressed in the study of
archaeological human remains? Bioarchaeologists are not merely counting the lesions on
skeletons and the shapes of skeletal elements in order to characterize variation. Rather, once-
living people resided within a complex web of circumstances involving social relations,
cultural behaviors, settlement systems, environmental conditions, and a range of other
interlinked contexts that we, as anthropologists, routinely take into account. In interpreting
health patterns, Goodman (1993) advocated use of multiple health indicators, rather than
single skeletal indicators. It is the record of these multiple indicators viewed within these
larger contextual domains that provides for a more informed understanding of the bioarch-
aeological record.
Bioarchaeology today is committed to the understanding of the social context and its role
in framing skeletal biology. Social contexts provide an important opportunity for interpreting
the range of variation in the behavioral record discussed throughout this book. The strength
of “social” bioarchaeology is especially well illustrated in the range of investigations
discussed, but one of the theoretical breakthroughs regarding the relationship between
old approaches and new ways of looking at variation as it pertains to biodistance

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 13:55:51, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.013
424 Bioarchaeology: skeletons in context

research. Anthropologists have had a centuries-old interest in historical issues – Where did a
specific population or society come from? Within a population, who is related to whom? These
are baseline questions, fueling not an inconsiderable interest in what we do by both the public
and scientists alike. Historically, these questions have been asked in the context of classifica-
tion and racial typology. Certainly, new methods of answering questions relating to identity
have been developed and may have their historical origins in the eighteenth and nineteenth
centuries, and result simply in another rendition of old and debunked notions about human
variation (Armelagos & Van Gerven, 2003; Armelagos et al., 1982). A new generation of
bioarchaeologists advocates the view that biodistance analysis offers an important opportun-
ity to address, in compelling ways, socially relevant problems having meaning for anthropol-
ogy in particular and social science in general (Stojanowski, 2013a). There are a number of
reasons why we do biodistance analysis in the twenty-first century (see Chapter 9), but among
the most important is the study of group identity, whether it is based on family, community,
ethnicity, or some other affiliative context. This ethnogenetic approach focuses on diachronic
processes of group emergence or reaffiliation. These processes are expressed synchronically
between individuals and social groups, such as between communities or tribes at any
particular time. These ethnogenetic events occur in many forms, including via intermarriage
between communities, large-scale migration and gene flow, and sometimes, by replacement
of entire populations. As is so well articulated by Stojanowski (2013a:72), “Bioarchaeology,
by virtue of the temporal resolution afforded archaeological assemblages, has the ability to
elucidate both synchronic and diachronic elements of ethnogenesis in past populations – the
former by considering patterns of genetic/phenetic variation against socially defined criteria
of group distinction; the latter by reconstructing patterns of gene flow among populations
through time.”

11.2 Framing the problems and questions: it is all about the hypothesis
Understanding data sets derived from the study of ancient skeletons and their broader
context is only one part of how best to approach the bioarchaeological record. Of equal
importance is the development of hypotheses that serve to frame the research. By construct-
ing research around a hypothesis, bioarchaeologists are able to develop a set of clearly
articulated questions and to generate expectations, to select skeletal series appropriate in
size and composition for the questions asked, to identify or develop the methods –
sometimes appropriately involving very simple methods, but sometimes involving complex
methods and supporting technology – to collect the data, analyze the data, and interpret the
results. Whether we arrive at the responses to our hypotheses through inductive or deductive
approaches, we must ask of our own research or of the research of others: Are the results
consistent with the original expectations that were established at the beginning of the
project, especially taking into account demographic composition, taphonomic history and
biases, and cultural, social, and behavioral factors that enter into who is or is not included in
mortuary assemblages?
There are a number of well-known examples of hypotheses tested in bioarchaeology. One
that has dominated much of my own research agenda pertaining to recent human evolution is
the foraging-to-farming transition, a global event beginning about 10–12 thousand years
ago. Working in a wide range of areas of the world in different geographic and temporal

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 13:55:51, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.013
11.2 Framing the problems and questions: it is all about the hypothesis 425

scales over the last half century, bioarchaeologists have undertaken independent regionally
based research programs in North and South America, Europe, Asia, and Africa that have
investigated health and behavioral outcomes in populations that underwent the transition
from foraging to farming in earlier prehistory or from less intensified to more intensified
farming in later prehistory (Cohen, 1989; Cohen & Armelagos, 1984; Cohen & Crane-Kramer,
2007; Eshed et al., 2010; Pechenkina & Oxenham, 2013; Pechenkina et al., 2013; Pinhasi &
Stock, 2011; Roberts & Cox, 2003; Steckel & Rose, 2002; Steckel et al., 2002; and see
summaries in Larsen, 1995, 2006; but see Douglas & Pietrusewsky, 2007; Oxenham, 2006;
Tayles et al., 2009 for an alternative perspective). All of this work addresses the hypothesis
that the foraging-to-farming transition led to an elevation in morbidity in particular and a
decline in the quality of life in general. These investigations showed a consistent pattern of
increasing morbidity, namely, higher prevalence of infectious lesions, enamel growth arrest,
reduced growth velocity, and other outcomes that all point to a decline in health. In a highly
consistent manner, these findings meet the expectation of negative health impacts involving a
shift from foraging and collecting to one where diet either shifted to or increased reliance on
domesticated cereals. This expectation is derived from the fact that domesticated cereals have
the following attributes: (1) they are deficient in one or more essential amino acids (e.g., lysine
for maize, millet, and wheat); (2) they lack of sufficient bioavailable iron; (3) they are deficient
in one or more vitamins; for example, B1 (thiamine), B2 (riboflavin), C (ascorbic acid); (4) they
have links with malnutrition, immunosuppression, and susceptibility to a variety of patho-
gens, rendering the individual prone to infection; and (5) they are carbohydrates, creating
circumstances that promote a cariogenic oral environment and dental caries, oral infection,
and susceptibility to degenerative conditions in later life (e.g., cardiovascular disease and
diabetes).
Within this larger question, one area of study that has demonstrated the consistency of
outcomes is skeletal growth. It is well understood that human populations experiencing
elevated stress are generally growth-retarded, with respect to both attainment per age in
juveniles and terminal height in adults (see Chapter 2). Temporal comparisons within arch-
aeological series reveal reduction in growth velocity and ultimately, adult height, and reflect
negative responses to elevated disease, physiological stress, and nutritional insult. Based on
comparison of attainment of growth in archaeological skeletons and in living populations, it
is difficult to tease apart the precise roles of nutrition and genetics, which are the primary
determinants of body size (Bogin, 1999). Thus, based on growth attainment per age, we do not
know precisely what factors are involved in its determination, or, for the purposes of this
discussion, whether growth attainment observed in skeletal series simply reflects a sample of
nonsurvivors who represent mortality bias rather than being an actual indication of stress
(Lovejoy et al., 1990; Saunders & Hoppa, 1993).
Saunders and Hoppa (1993) addressed this conundrum via their comparison of survivors
and nonsurvivors in living populations and their discussion of the implications for past
populations. They observed that nonsurvivors in living populations have higher morbidity
per age than survivors; nonsurvivors may have had high levels of stress, indeed resulting in
shorter height-for-age than survivors. Modeling the magnitude of mortality bias in growth by
projecting various survivor versus nonsurvivor distributions of height-for-age in living
populations shows that survivors are taller for age than nonsurvivors. This is consistent
with expectations, as it is important to realize that nonsurvivors are not selected against

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 13:55:51, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.013
426 Bioarchaeology: skeletons in context

simply because they are short. Rather, they are selected against because of the suite of
biological stressors (e.g., disease, undernutrition) that contributed to their reduced height-
for-age in the first place. However, simple comparisons of long bone length at various ages
reveal minimal differences, below 3 mm for femur length at 12 years of age (Saunders &
Hoppa, 1993: Table 4). These findings strongly suggest that the biological mortality bias
observed in long bone length of juvenile skeletons from archaeological settings is minimal.
Other factors far outweigh growth differences that might be present between survivors and
nonsurvivors, including bias in age estimation techniques, sample size, and preservation
status (Saunders & Hoppa, 1993).
There are fundamental differences in growth between human populations that are reflected
in variation in developmental patterns of long bones. Comparisons of the prehistoric Libben
series with modern populations from Denver, Colorado, reveal potentially large growth
differences (attainment per age) between homologous skeletal elements (Lovejoy et al.,
1990). In view of the fact that the Libben and Denver populations are from very different
environmental and sociocultural settings, these differences are to be expected. The pattern of
growth during the postnatal period between the two groups is remarkably similar, despite high
stress levels due to elevated systemic infection in the Libben series (Saunders et al., 1995).
These findings concur with Saunders and Hoppa’s (1993) conclusion that mortality bias is not
a significant deterrent to the study of growth in past human groups.
Other comparative analyses of body mass from juvenile remains based on articular dimen-
sions of lower limb bones (Ruff et al., 2013; Sciulli & Blatt, 2008) and midshaft femoral polar
second moments of area (Robbins et al., 2010) also match expectations in comparison with
modern populations. In addition, estimates of body mass and stature in juveniles from
historic-era Arikara from the Sully site, South Dakota, Neolithic Çatalhöyük, Turkey, and
twentieth-century Americans (Denver, Colorado) show similarities and differences that are
important for interpreting variation in health and well-being in these samples. In particular,
Arikara children had slower growth than Çatalhöyük and modern children in earlier age
groups (Ruff et al., 2013). These findings match the record from evidence of relatively poor
health in historic-era Arikara (Jantz & Owsley, 1984a) and relatively better health in Denver
and Çatalhöyük children (Hillson et al., 2013).
My point here is not that the growth record (and health generally) in recent Homo sapiens
either deteriorated or improved with agriculture or that historic-era Arikara were relatively
more worse off than Neolithic Çatalhöyük inhabitants. Rather, the consistency of patterns
observed by various researchers suggests that selectivity has not appreciably hampered
bioarchaeologists’ efforts to evaluate and characterize health in the past. Nevertheless,
researchers must be diligent in their evaluations of skeletal series by drawing on a range
of contextual observations, including cultural and social setting, multiple skeletal and
dental stress indicators, and health status in contemporary settings that might inform our
understanding of past groups (Wood et al., 1992; and see Goodman, 1993, and Wood &
Milner, 1994).
In summary, this book promotes a science-based approach to bioarchaeology. In the
absence of testable hypotheses, studies of skeletons or application of technology, simple or
complex, are mere description. Rather, the construction of bioarchaeological research should
be informed by hypotheses and questions that address fundamental issues about human
behavior viewed broadly.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 13:55:51, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.013
11.2 Framing the problems and questions: it is all about the hypothesis 427

Bioarchaeology as a modern anthropological science puts us in a very special position to


characterize and understand the origins, variation, and patterns of modern human
conditions – social, cultural, and biological – in a highly integrative way, more so than
intended when the field began to shift from a largely descriptive endeavor to one focusing on
process. The successes in this regard stem from the solid foundation upon which the field is
built: a population-oriented perspective, the fundamental role of culture as an adaptive
element linked with biological adaptation, the availability of long chronologies for assessing
change, and the power of the scientific method involving testing of hypotheses and generat-
ing questions about the human experience (Zuckerman & Armelagos, 2011). The field has
become global with remarkable temporal and geographic perspective that represents the best
of the vision that anthropology as a discipline shares broadly. Bioarchaeology, in particular,
has been highly successful in integrating itself with other sciences in an interdisciplinary
fashion. This is readily apparent from a number of contexts discussed in this book, ranging
across aDNA analysis for characterizing ancient pathogens and developing insights into
pathogen evolution and human disease history (Wilbur & Stone, 2012), use of microcomputed
tomography and other imaging systems for documenting bone cross-sections and ontogenetic
changes in activity and behavior as they relate to workload in ancient populations (Gosman
et al., 2013; Larsen et al., 2013), experimental approaches to understanding skeletal morph-
ology and behavior in living humans (Macdonald et al., 2009; Schmitt et al., 2003), ethno-
graphic context and interpreting patterns of oral health (Harrod et al., 2012; Walker &
Hewlett, 1990; Walker et al., 1998), and the remarkable range of issues beyond dietary
reconstruction, including mobility, life history, social stratification, household and social
identity, and health and well-being made possible by analysis of a number of stable isotopes
(Bentley et al., 2005; Knudson, Pestle et al., 2012; Pearson, 2013; Pilloud & Larsen, 2011;
Reitsema, 2013; Richards & Montgomery, 2012; Tung & Knudson, 2008). These are technol-
ogies that previous generations of bioarchaeologists could not have envisioned possible. It is,
however, important to reiterate that these technologies do not make bioarchaeological
research important to modern science. Rather, it is the opportunities that this technology
provides in addressing pressing problems, issues, and hypotheses about the past and present
world that makes it so important and exciting.
Three decades ago, Armelagos and others (Armelagos et al., 1982; Lovejoy et al., 1982)
raised the concern that technology simply provided a new venue for the continuation of
typological research and that the study of human remains was largely descriptive. They
concluded that rather than the problem directing the study of human remains from archaeo-
logical settings, it is the data that are directing the study. At the time they were writing in the
early 1980s, the field was still submerged in a highly descriptive past. Nevertheless, theoretical
issues were not lacking, including effects of health on major demographic transitions, the
impact of diet on craniofacial and postcranial phenotypic variation, and the foraging-to-
farming transition. Simply, answers without questions or hypotheses are not moving the field
forward. The good news is the fact that the field is advancing. This progress stems from the
fact that if we are to make the research predictive and inferential, then the products of
bioarchaeological research will move beyond the typological past (Larsen, 2010; for other
viewpoints, see Armelagos & Van Gerven 2003; Stojanowski & Buikstra, 2005). Ultimately,
the skeletons hold the answers. Our goal is to frame the important questions that will result in
meaningful new knowledge.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 13:55:51, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.013
428 Bioarchaeology: skeletons in context

11.3 Ethics in bioarchaeology


The remains of the dead derived from past societies have very real ties to the present-day
world. Virtually every living society venerates their dead. Members of these societies identify
themselves as descendants of the deceased, often with strongly perceived social bonds.
Around the world, descendants of pre-colonial, indigenous communities have argued that
because they have a relationship with once-living groups now represented by skeletal
remains, the disposition of bones and teeth should be controlled exclusively by them. In
Tasmania and Australia, backed by stringent cultural heritage laws, Aboriginal groups have
been able to assert their claims to remains for repatriation and reburial (Chari & Lavallee,
2013; Howitt, 1998; Morell, 1995; Walker, 2004). In Israel, new, more conservative interpret-
ations of the 1978 antiquities law have virtually halted research on archaeological human
remains younger than 5000 years old. In South Africa, the return of the remains of native
South Africans has promoted a debate on the ethics of scientific study of human remains
(Morris, 2009; Steyn & Nienaber, 2005).
In the United States, concerns regarding the control of native human remains by Native
Americans was first expressed in the early 1960s (Anderson et al., 1978). Within a span of a
few years, these concerns spread rapidly among various tribes. Opinions about the disposition
of skeletons from archaeological sites currently housed in museum collections vary widely
among native groups (Kakaliouras, 2012; Mihesuah, 2000; Ubelaker & Grant, 1989). Never-
theless, the consensus among native groups was well in place regarding the shift in control of
skeletal collections from museums and other institutions to tribes, resulting in the develop-
ment of regulations by most states for excavation and disposition of native human remains
(Ousley et al., 2005; reviewed by Ubelaker & Grant, 1989). In 1990, the Native American
Graves Protection and Repatriation Act (NAGPRA; US Public Law 101–601) was enacted,
mandating that all US Government agencies, non-Smithsonian Institution museums
(P.L. 101–185 applies to the Smithsonian Institution), and other institutions receiving federal
funding must inventory Native American human remains, assess ancestral associations
(“cultural affiliation”), communicate with federally recognized tribes, and return remains if
requested by these tribes. Additional NAGPRA regulations addressing the disposition of
unaffiliated remains went into effect in 2010 (Department of Interior, 2010).
The NAGPRA legislation and regulations of 1990 and 2010 have had important implica-
tions for curation and the study of Native American human remains in the United States. As
noted by William Fitzhugh, “. . .the sacrosanct principle of protecting valuable scientific
collections was replaced by a new set of untested and vaguely defined guidelines. For the
Native Americans, the case was a turning point in a long, painful history of invasive studies
conducted by an often insensitive establishment” (1994:vii). Owing to these demands by
native groups, culturally affiliated remains have a different future than the one envisioned
in the mid-twentieth century. Some interactions between the scientific and Native American
communities have been negative since the passing of NAGPRA or similar agreements between
states or other official bodies and native groups. However, in many settings, these develop-
ments have opened up channels of communication and collaboration between native peoples
and anthropologists in an unprecedented fashion as well as facilitating the development of
clearly articulated ethical practices (Kakaliouras, 2008, 2012; Lambert, 2012; Larsen &
Walker, 2005; McGuire, 1994; Walker, 2008). Far from spelling doom for bioarchaeology,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 13:55:51, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.013
11.4 Bioarchaeology looking forward 429

the dialog is resulting in a more informed understanding of native groups, developing an


appreciation for their concerns, and working collectively to understand their ancestors.
Undoubtedly, human remains will be repatriated and reburied; data will be lost. However, it
is essential that (1) the scientific community respect the perspectives and desires of living
descendants, (2) descendants have the authority to control the remains of deceased relatives,
and (3) when possible, human remains may be preserved for continued study and increased
understanding of the shared human past. Rather than ending bioarchaeology in the United
States, the climate of research wrought by NAGPRA, requiring communication between
scientists and native groups, has presented an important opportunity to inform native groups
and the public about the centrality of human remains for interpreting the past. The chance is
now in hand for sharing this information widely, especially regarding the large and crucial
role for human biology and bioarchaeology in understanding the history of the human
condition. It is simply no longer the case that there is one community that has privileged
access to remains from archaeological contexts. Rather, the other stakeholders must include
descendant communities having important links with the past (Blakey, 2010).

11.4 Bioarchaeology looking forward


At no other time in the history of bioarchaeology has there been so much public and scientific
interest in the discipline. While some public interest may be piqued by reports on something
characterized by the popular media as gruesome or odd, usually associated with the death of a
famous personage in some dramatic fashion, the underlying story is usually based in the
domain of science and application of new and evolving methods that address a specific
hypothesis or question about the past. Probably the most cited are the advances in molecular
biology and especially analysis of ancient DNA in the study of pathogens in the past.
New amplification methods and controls for contamination are making it possible to study
the evolution of specific pathogens, such as Mycobacterium tuberculosis and M. leprae
(Schuenemann et al., 2013; Wilbur & Stone, 2012). This work will likely soon provide new
perspectives on pathogens, parasitic diseases, genetic diseases, and susceptibility to infection.
Importantly, it will also provide an important context for understanding origins of conditions
that continue to plague societies around the world today. Equally remarkable is the study of
the evolution of bacterial pathogens and co-evolution with key events in recent human
health. Genomic research on DNA of Streptococcus mutans reveals that an exponential
increase in strains of S. mutans began about 10 000 yBP, likely at the same time that humans
first began to cultivate plants (Cornejo et al., 2013). This suggests that most of the genes found
in S. mutans are associated with metabolic processes involving sugar/starch metabolism or
acid tolerance, the environment matching the human mouth. That is, S. mutans flourishes in
cariogenic plaque, a factor made possible by its ability to grow, reproduce, and digest
carbohydrates in a relatively low pH setting (Cornejo et al., 2013). While still open to debate,
this work suggests that the foraging-to-farming transition likely formed new associations
between human populations and specific oral pathogens. Such work has important links with
bioarchaeology, especially in light of advances made in the identification of pathogenic
bacteria in calculus in archaeological teeth (Adler et al., 2013; de la Fuente et al., 2013; Preus
et al., 2011; Warinner et al., 2014). Indeed, analysis of calculus from pre-farming Mesolithic,
Neolithic, Bronze Age, late Medieval, and nineteenth-century individuals from Europe reveals

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 13:55:51, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.013
430 Bioarchaeology: skeletons in context

much greater prevalence of S. mutans and lower microbial diversity in modern Europeans
than in early foragers and the first farmers (Adler et al., 2013). The reduced diversity of
microbiota, but including significant S. mutans, complements genomic analysis showing
elevation in particular strains of pathogenic oral microbiota. These important breakthroughs
underscore the relationship between carbohydrate consumption – a cultural event – and the
ecology of the human mouth. Importantly, the discovery reveals the challenges of oral health
in the world today, all within the context of the co-evolution of the human host and it
pathogens.
Other significant directions pertain to the use of experimental data in interpreting skeletal
morphology and its relationship with activity. Pioneered by Sherwood Washburn (1953) as an
element of the “New Physical Anthropology,” experimental approaches provide important
perspectives on skeletal morphology in functional context. As discussed in Chapter 6, there is
a growing understanding that populations pursuing a lifestyle with high levels of mobility
have distinctive signatures of diaphyseal morphology, which involved a general anterio-
posterior elongation of the femoral midshaft diaphysis, which is also revealed in distinctive
ratios of Ix/Iy or Imax/Imin (second moments of area about the x- and y-axes). Ruff and others
have shown that archaeological series comprising individuals having active, mobile lifestyles
have higher Ix/Iy or Imax/Imin ratios than do less active, sedentary populations (see Chapter 6).
Assessment of loading histories via computed tomographic analysis of midshaft tibiae of male
university cross-country runners, field hockey players, and non-athletes revealed that the
runners and hockey players have significantly higher values of areas and second moments of
area (J, %CA, Imax, and Imin) than the non-athletes (Shaw & Stock, 2009b; and see Macdonald
et al., 2009). These and other experimentally based analyses of behavioral outcomes are a
direction from which bioarchaeology will benefit (Schmitt & Churchill, 2003; Teaford & Lytle,
1996).
Similarly, the importance of collaborative study involving ethnographers and bioarch-
aeologists can be especially informative from the perspective of behavior. For example,
ethnobioarchaeological research on modern pygmy groups (Aka, Mbuti, and Efe) in the
Central African Republic revealed remarkable differences in access to food and nutritional
quality in regard to both gender and status, a finding that runs counter to the long-held
assumption that pygmies are emblematic of egalitarian societies (Walker & Hewlett,
1990). Moreover, the ethnobioarchaeological study of dental caries and carbohydrate
consumption showed a clear link between the amount of carbohydrates consumed and
caries rates, but oral hygiene and differences in food preparation practices (e.g., cleaning
of eating utensils) also play an important part in oral health outcomes. Walker and
collaborators’ (1998) ethnobioarchaeological study of diet and dental caries among trad-
itional foragers/agriculturalists – Yanomami (Venezuela), Yora (Peru), and Shiwiar
(Ecuador) – revealed that the relationship between oral health (especially dental caries)
and food consumption practices is complex. For example, the Yanomami have lower
caries rates than the Yora, even though the Yanomami consume less meat than the Yora.
Unlike the Central African pygmies, there are quite low differences between men and
women, despite the more frequent presence of manioc in womens’ mouths as they chew it
for beer production. However, the constant movement of manioc in the mouth and
frequent spitting appear to reduce the cariogenic effects of the food. These investigations

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 13:55:51, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.013
11.4 Bioarchaeology looking forward 431

are illuminating because they show the complex nature of food consumption and tech-
nology, and by doing so, they provide key insights into behavioral reconstructions. What
is especially important about this work is that it controls for variables observable in the
living that may be applied through carefully constructed analogies to the study of skeletal
remains.
The aforementioned research also speaks to the profound importance of understanding
the inextricable links between biology, culture, and society. A large and impressive record
involving biocultural perspectives on the study of human remains from archaeological
contexts has revolutionized the field, promoting the transition from what was once
largely a descriptive enterprise to one that is informed by social, economic, cultural,
and political circumstances of once-living societies. Importantly, this work on social and
behavioral circumstances in past populations can be used to document the circumstances
of health disparities. Similarly, circumstances resulting in health disparities in the living
form an experimental platform for interpreting the record of past health and well-being
(Goodman et al., 1992; Harrod et al., 2012; May et al., 1993; Walker & Hewlett, 1990;
Walker et al., 1998).
Among the most promising areas of bioarchaeology moving forward is the study of
changing social relations, and especially the essential role that violence has played in past
societies and at all levels, from interpersonal relations to developments as complex as state
formation (various in Martin et al., 2012; Tung, 2012) and colonization of New World
native societies (Klaus, 2012; Klaus et al., 2009; Stojanowski, 2009, 2013b). In addition to
these developing and emerging frontiers, the field is making tremendous strides in moving
away from one that is focused solely on understanding the past to one that is engaged with
and socially relevant to the present. It is doing this by providing a large and robust record
of the history of the human condition, including especially key aspects of health and well-
being.
This is important because it provides a picture of the human experience laying the
foundation for understanding today’s world. That is, the study of the evolution of pathogens
provides answers to questions about key diseases today, including those now re-emerging
(e.g., tuberculosis). It also provides a means for testing hypotheses about health in early life
and its impact on later life. There is a growing record that early life experiences (e.g., low birth
weight and various forms of fetal stress) predict elevated prevalence of degenerative diseases,
such as cardiovascular disease and diabetes in adulthood. In this regard, the bioarchaeological
record provides evidence of morbidity during the period of the growing dentition. Analysis of
enamel defects as a record of growth arrest in response to environmental perturbations
demonstrates a link between early life stresses and increased mortality (Armelagos et al.
2009).
Bioarchaeology holds considerable promise for addressing other issues relating to current
and ongoing concerns about climate change and its role in health outcomes. Clearly,
climate has played an essential role in explaining at least some of the variation in distribution
of specific diseases in the past and today (Drake & Oxenham, 2012; McMichael, 2010;
Patz et al., 2002; Stewart, 1960). Although climate change may not affect health
patterns directly, the alteration in foods available and loss of resources are important elements
for interpreting dietary and health change. For example, bioarchaeological research in the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 13:55:51, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.013
432 Bioarchaeology: skeletons in context

American Great Basin and west-central India reveal changing climate that resulted in a
pattern of increasing aridity and abandonment of agriculture. Both settings show clear
patterns of health and lifestyle change associated with major climatic events, to include
altered growth patterns, mobility and workload, and living circumstances overall (Coltrain &
Leavitt, 2002; Coltrain & Stafford, 1999; Robbins Schug, 2011b; Ruff, 1999; Temple, 2008).
There is a clear opportunity in hand to apply bioarchaeological inquiry and understanding of
the fundamental dynamics of the past in a way that powerfully informs the present and
future.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 21 Jan 2017 at 13:55:51, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.013
REFERENCES

Every effort has been made to secure necessary permissions to reproduce copyright material in this work,
although in some cases, it has proved impossible to trace copyright holders. If any omissions are brought
to our notice, we will be happy to include appropriate acknowledgments on reprinting [in any
subsequent edition].

Abbate LM, Stevens J, Schwartz TA, et al. 2006. shows changes in oral microbiota with dietary
Anthropometric measures, body composition, shifts of the Neolithic and Industrial
body fat distribution, and knee osteoarthritis. revolutions. Nature Genetics 45:450–456.
Obesity 14:1274–1281. Adler CJ, Haak W, Donlon D, Cooper A, The
Abbott S, Trinkaus E, Burr DB. 1996. Dynamic Geographic Consortium. 2011. Survival and
bone remodeling in later Pleistocene fossil recovery of DNA from ancient teeth and bones.
hominids. American Journal of Physical Journal of Archaeological Science 38:956–964.
Anthropology 99:585–601. Agarwal SC. 2008. Light and broken bones:
Abdushelishvili MG. 1984. Craniometry of the examining and interpreting bone loss and
Caucasus in the Feudal period. Current osteoporosis in past populations. In:
Anthropology 25:505–509. Katzenberg MA, Saunders SR, editors.
Abram AC, Keller TS, Spengler DM. 1988. The Biological Anthropology of the Human
effects of simulated weightlessness on bone Skeleton, Second Edition. Hoboken, NJ:
biomechanical and biochemical properties in Wiley-Liss; p. 387–410.
the maturing rat. Journal of Biomechanics Agarwal SC, Beauchesne P. 2011. It is not carved
21:755–767. in bone: development and plasticity of the aged
Acheson RM. 1959. Effects of starvation, skeleton. In: Agarwal SC, Glencross BA,
septicaemia and chronic illness on the growth editors. Social Bioarchaeology. Chichester, UK:
cartilage plate and metaphysis of the immature Wiley-Blackwell; p. 314–334.
rat. Journal of Anatomy 93:123–130. Agarwal SC, Dumitriu M, Tomlinson GA, Grynpas
Ackerknecht EH. 1967. Primitive surgery. In: MD. 2004. Medieval trabecular bone
Brothwell D, Sandison AT, editors. Diseases in architecture: the influence of age, sex, and
Antiquity. Springfield, IL: Charles C. Thomas; lifestyle. American Journal of Physical
p. 635–650. Anthropology 124:33–44.
Acsádi G, Nemeskéri J. 1970. History of the Agarwal SC, Glencross B. 2011. Building a social
Human Life Span and Mortality. Budapest, bioarchaeology. In: Agarwal SC, Glencross B,
Hungary: Akadémiai Kiadó. editors. Social Bioarchaeology. Hoboken, NJ:
Adachi N, Shinoda K-I, Umetsu K, Matsumura H. Wiley-Blackwell; p. 1–11.
2009. Mitchondrial DNA analysis of Jomon Agarwal SC, Stout SD, editors. 2003. Bone Loss and
skeletons from the Funadomari site, Hokkaido, Osteoporosis: An Anthropological Perspective.
and its implications for the origins of Native New York, NY: Kluwer Academic Plenum.
American. American Journal of Physical Agnew AM, Bolte JH IV. 2012. Bone fracture:
Anthropology 138:255–265. biomechanics and risk. In: Crowder C, Stout S,
Adams P. 1969. The effect of experimental editors. Bone Histology: An Anthropological
malnutrition on the development of long Perspective. Boca Raton, FL: CRC Press;
bones. Bibliotheca Nutritio et Dieta 13:69–73. p. 221–240.
Adler CJ, Dobney K, Weyrich LS, et al. 2013. Ahlborg H, Johnell O, Turner C, Rannevik G,
Sequencing ancient calcified dental plaque Karlsson MK. 2003. Bone loss and bone size

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
434 References

after menopause. New England Journal of Allen KD. 2010. Racial and ethnic disparities in
Medicine 349:327–334. osteoarthritis phenotypes. Current Opinion in
Aikens CM, Higuchi T. 1982. Prehistory of Japan. Rheumatology 22:528–532.
New York, NY: Academic Press. Allen T, Novak SA, Bench LL. 2007. Patterns of
Ajwani S, Mattila KJ, Narhi TO, Tilvis RS, injuries: accident or abuse. Violence Against
Ainamo A. 2003. Oral health status, C-reactive Women 13:802–816.
protein and mortality – a 10 year follow-up Allen WH, Merbs CF, Birkby WH. 1985. Evidence
study. Gerodontology 20:32–40. for prehistoric scalping at Nuvakwewtaqa
Akins NJ. 1986. A Biocultural Approach to (Chavez Pass) and Grasshopper Ruin, Arizona.
Human Burials from Chaco Canyon, New In: Merbs CF, Miller RJ, editors. Health and
Mexico. National Park Service, Reports of the Disease in the Prehistoric Southwest. Arizona
Chaco Center, No. 9. State University. Anthropological Research
Albert AM, Greene DL. 1999. Bilateral asymmetry in Papers, No. 34; p. 23–42.
skeletal growth and maturation as an indicator Allison MJ. 1984. Paleopathology in Peruvian and
of environmental stress. American Journal of Chilean populations. In: Cohen MN, Armelagos
Physical Anthropology 110:341–350. GJ, editors. Paleopathology at the Origins of
Albert AM, Ricanek K Jr., Patterson E. 2007. Agriculture. Orlando, FL: Academic Press;
A review of the literature on the aging adult p. 515–529.
skull and face: implications for forensic science Allison MJ, Gerszten E, Munizaga J, Santoro C,
research and applications. Forensic Science Mendoza D. 1981. Tuberculosis in Pre-
International 172:1–9. Columbian Andean populations. In: Buikstra
Alexandersen V. 1967. The pathology of the jaws JE, editor. Prehistoric Tuberculosis in the
and temporomandibular joint. In: Brothwell D, Americas. Center for American Archeology,
Sandison AT, editors. Diseases in Antiquity. Scientific Papers, No. 5; p. 49–61.
Springfield, IL: Charles C. Thomas; p. 551–595. Allison MJ, Mendoza D, Pezzia A. 1973.
Alexandersen V. 1988. The late-Mesolithic Documentation of a case of tuberculosis in
dentition in southern Scandinavia. Rivista di Pre-Columbian America. American Review
Antropologia (Roma) 66(S):191–204. of Respiratory Diseases 107:985–991.
Alexandre C, Vico L. 2011. Pathophysiology of Al-Shammery A, el Backly M, Guile EE. 1998.
bone loss in disuse osteoporosis. Joint Bone Permanent tooth loss among adults and
Spine 78:572–576. children in Saudi Arabia. Community Dental
Alfonso MP, Standen VG, Castro MV. 2007. The Health 15:277–280.
adoption of agriculture among northern Chile Alt KW, Benz M, Muller W, et al. 2013. Earliest
populations in the Azapa valley, 9000–1000 evidence of social endogamy in the 9,000-
BP. In: Cohen MN, Crane-Kramer MM, editors. year-old population of Basta, Jordan. PLoS
Ancient Health: Skeletal Indicators of ONE 8(6)e65649:1–9.
Agricultural and Economic Intensification. Alt KW, Vach W. 1998. Kinship studies in skeletal
Gainesville, FL: University Press of Florida; remains – concepts and examples. In: Alt K,
p. 113–129. Rösing FW, Teschler-Nicola M, editors. Dental
Alfonso-Durruty MP. 2011. Experimental Anthropology: Fundamentals, Limits, and
assessment of nutrition and bone growth’s Prospects. Vienna, Austria: Springer-Verlag;
velocity effects on Harris lines formation. p. 537–554.
American Journal of Physical Anthropology Alvarez JO. 1995. Nutrition, tooth development,
145:169–180. and dental caries. American Journal of Clinical
Allander E. 1995. Epidemiology of osteoporosis. Nutrition 61:410S–416S.
In: Anderson JJB, Garner SC, editors. Calcium Alvarez JO, Lewis CA, Saman C, et al. 1988.
and Phosphorus in Health and Disease. Boca Chronic malnutrition, dental caries, and tooth
Raton, FL: CRC Press; p. 355–362. exfoliation in Peruvian children aged 3–9

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 435

years. American Journal of Clinical Nutrition carbon and nitrogen isotope ratios. Nature
48:368–372. 319:321–324.
Alvarez JO, Navia JM. 1989. Nutritional status, Ambrose SH, Krigbaum J. 2003. Bone chemistry
tooth eruption, and dental caries: a review. and bioarchaeology. Journal of
American Journal of Clinical Nutrition Anthropological Archaeology 22:193–199.
49:417–426. Ambrose SH, Norr L. 1992. On stable isotopic data
Alvrus A. 1999. Fracture patterns among the and prehistoric subsistence in the Soconusco
Nubians of Semna South, Sudanese Nubia. region. Current Anthropology 33:401–404.
International Journal of Osteoarchaeology Ambrose SH, Norr L. 1993. Isotopic composition of
9:417–429. dietary protein and energy versus bone
Alzualde A, Izagirre N, Alonso S, et al. 2007. collagen and apatite: purified diet growth
Influences of the European kingdoms of Late experiments. In: Lambert JB, Grupe G, editors.
Antiquity on the Basque Country. Current Prehistoric Human Bone: Archaeology at the
Anthropology 48:155–163. Molecular Level. Berlin, Germany: Springer-
Ambrose SH. 1987. Chemical and isotopic Verlag; p. 1–37.
techniques of diet reconstruction in eastern Andersen JG. 1969. Studies in Medieval diagnosis
North America. In: Keegan WF, editor. of leprosy in Denmark: an osteoarchaeological,
Emergent Horticultural Economies of the historical and clinical study. Danish Medical
Eastern Woodlands. Southern Illinois Bulletin 16(S9):1–142.
University at Carbondale, Center for Anderson BG, Stevenson PH. 1930. The occurrence
Archaeological Investigations, Occasional of mottled enamel among the Chinese. Journal
Paper, No. 7; p. 78–107. of Dental Research 10:233–238.
Ambrose SH. 1990. Preparation and Anderson DD, Finnegan M, Hottop J, Fisher A.
characterization of bone and tooth collagen for 1978. The Lewis Central School site (13PW5): a
isotopic analysis. Journal of Archaeological resolution of ideological conflicts at an Archaic
Science 17:431–451. ossuary in western Iowa. Plains Anthropologist
Ambrose SH. 1993. Isotopic analysis of 23:193–219.
paleodiets: methodological and interpretive Anderson DG. 1994. The Savannah River
considerations. In: Sandford MK, editor. Chiefdoms: Political Change in the Late
Investigations of Ancient Human Tissue: Prehistoric Southeast. Tuscaloosa, AL:
Chemical Analyses in Anthropology. University of Alabama Press.
Langhorne, PA: Gordon and Breach Scientific Anderson JAD, Duthie JJR, Moody BP. 1962.
Publishers; p. 59–130. Social and economic effects of rheumatic
Ambrose SH, Buikstra J, Krueger HW. 2003. Status diseases in a mining population. Annals of the
and gender differences in diet at Mound 72 Rheumatic Diseases 21:342–352.
Cahokia, revealed by isotopic analysis of bone. Anderson JE. 1964. The People of Fairty. National
Journal of Anthropological Archaeology Museum of Canada Bulletin, No. 193.
22:217–226. Anderson JE. 1965. Human skeletons from
Ambrose SH, Butler BM, Hanson DB, Hunter- Tehuacan. Science 148:496–497.
Anderson RI, Krueger HW. 1997. Stable isotope Anderson JE. 1967. The human skeletons. In:
analysis of human diet in the Marianas Byers DS, editor. Prehistory of the Tehuacan
Archipelago, western Pacific. American Journal Valley, Volume 1. Environment and
of Physical Anthropology 104:343–361. Subsistence. Austin, TX: University of Texas
Ambrose SH, DeNiro MJ. 1986a. The isotope Press; p. 91–113.
ecology of East African animals. Oecologia Anderson JJB. 1995. Development and
69:396–406. maintenance of bone mass through the life
Ambrose SH, DeNiro MJ. 1986b. Reconstruction of cycle. In: Anderson JJB, Garner SC, editors.
African human diet using bone collagen Calcium and Phosphorus in Health and

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
436 References

Disease. Boca Raton, FL: CRC Press; Angel JL. 1966b. Early Skeletons from
p. 265–288. Tranquillity, California. Smithsonian
Anderson JJB, Pollitzer WS. 1994. Ethnic and Contributions to Anthropology, No. 2(1).
genetic differences in susceptibility to Angel JL. 1967. Porotic hyperostosis or
osteoporotic fractures. In: Draper HH, editor. osteoporosis symetrica. In: Brothwell D,
Advances in Nutritional Research, Volume 9. Sandison AT, editors. Diseases in Antiquity.
New York, NY: Plenum Press; p. 129–149. Springfield, IL: Charles C Thomas; p. 378–389.
Anderson JY, Trinkaus E. 1998. Patterns of sexual, Angel JL. 1968. Ecological aspects of
bilateral and interpopulational variation in palaeodemography. In: Brothwell D, editor. The
human neck-shaft angles. Journal of Anatomy Skeletal Biology of Earlier Human Population.
192:279–285. Oxford, UK: Oxford University Press;
Anderson T. 2001. Two decapitations from Roman p. 263–271.
Towcester. International Journal of Angel JL. 1969. The bases of paleodemography.
Osteoarchaeology 11:400–405. American Journal of Physical Anthropology
Andrushko VA, Latham KAS, Grady DL, Pastron 30:427–437.
AG, Walker PL. 2005. Bioarchaeological Angel JL. 1971a. The People of Lerna: Analysis of
evidence for trophy-taking in prehistoric a Prehistoric Aegean Population. Washington,
central California. American Journal of DC: Smithsonian Institution Press.
Physical Anthropology 127:375–384. Angel JL. 1971b. Early Neolithic skeletons from
Andrushko VA, Schwitalla AW, Walker PL. 2010. Çatal Hüyük: demography and pathology.
Trophy-taking and dismemberment as warfare Anatolian Studies 21:77–98.
strategies in prehistoric central California. Angel JL. 1972. Ecology and population in the
American Journal of Physical Anthropology eastern Mediterranean. World Archaeology
141:83–96. 4:88–105.
Andrushko VA, Torres EC. 2011. Skeletal evidence Angel JL. 1974. Patterns of fractures from
for warfare from the Cuzco region of Peru. Neolithic to modern times. Anthropologiai
American Journal of Physical Anthropology Közlemenyek 18:9–18.
146:361–372. Angel JL. 1975. Paleoecology, paleodemography
Andrushko VA, Verano JW. 2008. Prehistoric and health. In: Polgar S, editor. Population,
trepanation in the Cuzco region of Peru: a view Ecology and Social Evolution. Aldine, IL:
into an ancient Andean practice. American Chicago; p. 167–190.
Journal of Physical Anthropology 137:4–13. Angel JL. 1976. Colonial to modern change in the
Angel JL. 1944. Greek teeth: ancient and modern. U.S.A. American Journal of Physical
Human Biology 16:283–297. Anthropology 45:723–736.
Angel JL. 1947. The length of life in ancient Angel JL. 1978a. Pelvic inlet form: a neglected
Greece. Journal of Gerontology 2:18–24. index of nutritional status. American Journal
Angel JL. 1951. Population size and of Physical Anthropology 48:378.
microevolution in Greece. Cold Spring Harbor Angel JL. 1978b. Porotic hyperostosis in the
Symposium in Quantitative Biology eastern Mediterranean. Medical College of
15:343–351. Virginia Quarterly 14:10–16.
Angel JL. 1952. The human remains from Hotu Angel JL. 1979. Osteoarthritis in prehistoric
Cave, Iran. Proceediings of the American Turkey and medieval Byzantium. In: Cockburn
Philosophical Society 96:258–269. E, Duncan H, Riddle JM, editors. Arthritis:
Angel JL. 1964. The reaction area of the femoral Modern Concepts and Ancient Evidence. Henry
neck. Clinical Orthopaedics 32:130–142. Ford Hospital Medical Journal 27; p. 38–43.
Angel JL. 1966a. Porotic hyperostosis, anemias, Angel JL. 1982. A new measure of growth
malarias, and marshes in the prehistoric efficiency: skull base height. American Journal
eastern Mediterranean. Science 153:760–763. of Physical Anthropology 58:297–305.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 437

Angel JL. 1984. Health as a crucial factor in the Arkush E, Tung TA. 2013. Patterns of war in the
changes from hunting to developed farming in Andes from the Archaic to the Late Horizon:
the eastern Mediterranean. In: Cohen MN, insights from settlement patterns and cranial
Armelagos GJ, editors. Paleopathology at the trauma. Journal of Archaeological Research
Origins of Agriculture. Orlando, FL: Academic 21:307–369.
Press; p. 51–73. Armelagos GJ. 1969. Disease in ancient Nubia.
Angel JL, Kelley JO, Parrington M, Pinter S. 1987. Science 163:255–259.
Life stresses of the free Black community as Armelagos GJ. 1990. Health and disease in
represented by the First African Baptist Church, prehistoric populations in transition. In:
Philadelphia, 1823–1841. American Journal of Swedlund AC, Armelagos GJ, editors. Disease
Physical Anthropology 74:213–229. in Populations in Transition. New York, NY:
Angel JL, Olney LM. 1981. Skull base height and Bergin and Garvey; p. 127–144.
pelvic inlet depth from prehistoric to modern Armelagos GJ. 2000. Take two beers and call me in
times. American Journal of Physical 1,600 years. Natural History 109:50–54.
Anthropology 54:197. Armelagos GJ. 2003. Bioarchaeology as
Angle EH. 1898. Treatment of Malocclusion of the Anthropology. Archaeological Papers of the
Teeth. Philadelphia, PA: S.S. White Dental American Anthropological Association, No. 13;
Manufacturing Company. p. 27–40.
Antoine D, Hillson S, Dean C. 2009. The Armelagos GJ. 2008. Devouring ourselves. In:
developmental clock of dental enamel: a test Nichols DL, Crown PL, editors. Social Violence
for the periodicity of prism cross-striations in in the Prehispanic American Southwest.
modern humans and an evaluation of the most Tucson, AZ: University of Arizona Press;
likely sources of error in histological studies of p. 216–227.
this kind. Journal of Anatomy 214:45–55. Armelagos GJ, Carlson DS, Van Gerven DP. 1982.
Antoniades L, MacGregor AJ, Andrew T, Spector The theoretical foundations and development
TD. 2003. Association of birth weight with of skeletal biology. In: Spencer F, editor.
osteoporosis and osteoarthritis in adult twins. A History of American Physical Anthropology:
Rheumatology 42:791–796. 1930–1980. New York, NY: Academic Press;
Antunes-Ferreira N, Matos VMJ, Santos, AL. 2014. p. 305–328.
Leprosy in individuals unearthed near the Armelagos, GJ, Dewey J. 1970. Evolutionary
Ermida de Santo André and leprosarium of response to human infectious disease.
Beja, Portugal. Anthropological Science Bioscience 20:271–275.
121:149–159. Armelagos GJ, Goodman AH, Harper KN, Blakey
Apps MVB, Hughes TE, Townsend GC. 2004. The ML. 2009. Enamel hypoplasia and early
effect of birthweight on tooth-size variability mortality. Evolutionary Anthropology 18:
in twins. Twin Research 7:415–420. 261–271.
Arafat A. 1974. Periodontal status during Armelagos GJ, Jacobs KH, Martin DL. 1981. Death
pregnancy. Journal of Periodontology 45: and demography in prehistoric Sudanese
641–643. Nubia. In: Humphreys SC, King H, editors.
Arcini C. 1999. Health and Disease in Early Lund. Mortality and Immortality: The Anthropology
Lund, Sweden: Lund University Medical and Archaeology of Death. London, UK:
Faculty. Academic Press; p. 33–57.
Arden N, Nevitt MC. 2006. Osteoarthritis: Armelagos GJ, Mielke JH, Owen KH, Van Gerven
epidemiology. Best Practice & Research DP. 1972. Bone growth and development in
Clinical Rheumatology 20:3–25. prehistoric populations from Sudanese Nubia.
Arkush E. 2011. Hillforts of the Ancient Andes: Journal of Human Evolution 1:89–119.
Colla Warfare, Society and Landscape. Armelagos GJ, Van Gerven DP. 2003. A century of
Gainesville, FL: University Press of Florida. skeletal biology and paleopathology: contrasts,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
438 References

contradictions, and conflicts. American Atchley WR. 1993. Genetic and developmental
Anthropologist 105:53–64. aspects of variability in the mammalian
Armelagos GJ, Van Gerven DP, Martin DL, mandible. In: Hanken J, Hall BK, editors. The
Huss-Ashmore R. 1984. Effects of nutritional Skull, Volume 1. Development. Chicago, IL:
change on the skeletal biology of northeast University of Chicago Press; p. 207–247.
African (Sudanese Nubia) populations. In: Auerbach BM, Ruff CB. 2006. Limb bone bilateral
Clark JD, Brandt SA, editors. From Hunters asymmetry: variability and commonality
to Farmers: The Causes and Consequences among modern humans. Journal of Human
of Food Production in Africa. Berkeley, CA: Evolution 50:203–218.
University of California Press; Aufderheide AC. 1989. Chemical analysis of
p. 132–146. skeletal remains. In: İşcan MY, Kennedy KAR,
Armendariz J, Irigari S, Etxeberria F. 1994. New editors. Reconstruction of Life from the
evidence of prehistoric arrow wounds in the Skeleton. New York, NY: Alan R. Liss;
Iberian peninsula. International Journal of p. 237–260.
Osteoarchaeology 4:215–222. Aufderheide AC. 2003. The Scientific Study of
Armstrong E. 1984. Comment. Current Mummies. Cambridge, UK: Cambridge
Anthropology 25:318–319. University Press.
Arnaud CD, Sanchez SD. 1990. The role of calcium Aufderheide AC, Angel JL, Kelley JO, et al. 1985.
in osteoporosis. Annual Review of Nutrition Lead in bone. III. Prediction of social correlates
10:397–414. from skeletal lead content in four Colonial
Arnold J. 1992. Complex hunter-gatherer-fishers American populations (Catoctin Furnace,
of prehistoric California: chiefs, specialists, and College Landing, Governor’s Land, and Irene
maritime adaptations of the Channel Islands. Mound). American Journal of Physical
American Antiquity 57:60–84. Anthropology 66:353–361.
Arriaza BT. 1995. Beyond Death: The Chinchorro Aufderheide AC, Neiman FD, Wittmers LE Jr.,
Mummies of Ancient Chile. Washington, DC: Rapp G. 1981. Skeletal lead content as an
Smithsonian Institution Press. indicator of lifetime lead ingestion and the
Arriaza BT. 1997. Spondylolysis in prehistoric social correlates in an archaeological
remains from Guam and its possible etiology. population. American Journal of Physical
American Journal of Physical Anthropology Anthropology 55:285–291.
104:393–397. Aufderheide AC, Rodríguez-Martín C, editors.
Arriaza BT, Salo W, Aufderheide AC, Holcomb TA. 1998. The Cambridge Encyclopedia of Human
1995. Pre-Columbian tuberculosis in northern Paleopathology. Cambridge, UK: Cambridge
Chile: molecular and skeletal evidence. University Press.
American Journal of Physical Anthropology Aufderheide AC, Tieszen LL, Alison MJ, Wallgren
98:37–45. J, Rapp GJ. 1988. Chemical reconstruction of
Artelius T, Svanberg F. 2005. Dealing with the components in complex diets: a pilot study. In:
Dead: Archaeological Perspectives on Kennedy BV, LeMoine GM, editors. Diet and
Prehistoric Scandinavian Burial Ritual. Subsistence: Current Archaeological
Stockholm, Sweden: National Heritage Board. Perspectives. Archaeological Association of the
Ashworth A, Milner PF, Waterlow JC, Walker RB. University of Calgary, Proceedings of the 19th
1973. Absorption of iron from maize (Zea Annual Chacmool Conference; p. 301–306.
maize L.) and soya beans (Glycine hispida Aufderheide AC, Wittmers LE Jr., Rapp G Jr.,
Max.) in Jamaican infants. British Journal of Wallgren J. 1988. Anthropological applications
Nutrition 29:269–278. of skeletal lead analysis. American
Asingh P, Lynnerup P. 2007. Grauballe Man: An Anthropologist 90:931–936.
Iron Age Bog Body Revisited. Aarhus, Denmark: Austin C, Smith TM, Bradman A, et al. 2013.
Aarhus University Press. Barium distributions in teeth reveal early-life

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 439

dietary transitions in primates. Nature flow in Europe. American Journal of Human


498:216–219. Genetics 62:488–492.
Avis V. 1959. The relation of the temporal muscle Barbujani G, Sokal RR. 1990. Zones of sharp
to the form of the coronoid process. American genetic change in Europe are also linguistic
Journal of Physical Anthropology 17:99–104. boundaries. Proceedings of the National
Avis V. 1961. The significance of the angle of the Academy of Sciences 87:1816–1819.
mandible: an experimental and comparative Barker DJP. 2001. Fetal Origins of Cardiovascular
study. American Journal of Physical and Lung Disease. Oxford, UK: Taylor &
Anthropology 19:55–61. Francis Books.
Baer MJ. 1956. Dimensional changes in the human Barker DJP. 2012. Developmental origins of
head and face in the third decade of life. chronic disease. Public Health 126:185–189.
American Journal of Physical Anthropology Barker DJP, Eriksson JG, Forsen T, Osmond C.
14:557–575. 2002. Fetal origins of adult disease: strength of
Bailey SM, Gershoff SN, McGandy RB, et al. 1984. effects and biological basis. International
A longitudinal study of growth and maturation Journal of Epidemiology 31:1235–1239.
in rural Thailand. Human Biology 56:539–557. Barmes DE. 1962. Dental Disease Patterns Related
Bailit HL, Niswander JD, MacLean CJ. 1968. The to Dietary Patterns in Primitive Peoples of the
relationship among several prenatal factors Territory of Papua and New Guinea. DDS
and variation in the permanent dentition in Dissertation, University of Queensland,
Japanese children. Growth 32:331–345. Brisbane, Australia.
Bailit HL, Workman PL, Niswander JD, MacLean Barmes DE. 1977. Epidemiology of dental diseases.
CJ. 1970. Dental asymmetry as an indicator of Journal of Clinical Periodontology 4:80–93.
genetic and environmental conditions in human Barnes I, Thomas MG. 2006. Evaluating bacterial
populations. Human Biology 42:626–638. pathogen DNA preservation in museum
Baker BJ. 1997. Contributions of biological osteological collections. Proceedings of the
anthropology to the understanding of ancient Royal Society B 273:645–653.
Nubians and Nubian societies. In: Lustig J, Baron I, Hummel S, Herrmann B. 1996.
editor. Anthropology and Egyptology: Mycobacterium tuberculosis complex DNA in
A Developing Dialogue. Monographs in ancient human bones. Journal of
Mediterranean Archaeology 8; p. 106–116. Archaeological Science 23:667–671.
Baker BJ, Armelagos GJ. 1988. The origin and Barondess DA, Sauer NJ. 1985. Human skeletal
antiquity of syphilis: paleopathological remains from 20IS46: a Late Archaic burial
diagnosis and interpretation. Current locale in eastern Michigan. Michigan
Anthropology 29:703–737. Archaeologist 31:82–96.
Baker BJ, Dupras TL, Tocheri MW, Wheeler SM. Barrett AR, Blakey ML. 2011. Life histories of
2005. The Osteology of Infants and Children. enslaved Africans in colonial New York: a
College Station, TX: Texas A & M University bioarchaeological study of the New York
Press. African Burial Ground. In: Agarwal SC,
Bales CW, Anderson JJB. 1995. Influence of Glencross BA, editors. Social Bioarchaeology.
nutritional factors on bone health of the elderly. New York, NY: Wiley-Blackwell; p. 214–253.
In: Anderson JJB, Garner SC, editors. Calcium Barrett CK, Guatelli-Steinberg D, Sciulli PW. 2012.
and Phosphorus in Health and Disease. Boca Revisiting dental fluctuating asymmetry in
Raton, FL: CRC Press; p. 319–337. Neandertals and modern humans. American
Bamforth DB. 1994. Indigenous people, Journal of Physical Anthropology 149:
indigenous violence: precontact warfare on the 193–204.
North American Great Plains. Man 29:95–115. Barrett MJ, Brown T. 1966. Eruption of deciduous
Barbujani G, Bertorelle G, Chikhi L. 1998. teeth in Australian Aborigines. Australian
Evidence for Paleolithic and Neolithic gene Dental Journal 11:43–50.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
440 References

Bartelink EJ, Andrushko VA, Bellifemine VI, ancient Sudanese Nubia (A.D. 350). Science
Nechayev I, Jurmain R. 2014. Violence and 209:1532–1534.
warfare in the prehistoric San Francisco Bay Basu MK, Dutta AN. 1963. Report on “prevalence
area: regional and temporal variation in of periodontal disease in the adult population
conflict. In: Knüsel C, Smith MJ, editors. The in Calcutta,” by Ramfjord’s technique. Journal
Routledge Handbook of the Bioarchaeology of of the All India Dental Association 35:187.
Human Conflict. London, UK: Routledge; Batrawi AM. 1946. The racial history of Egypt and
p. 285–307. Nubia, Part II. Journal of the Royal
Barton L, Newsome SD, Chen F-H, et al. 2009. Anthropological Institute 76:132–156.
Agricultural origins and the isotopic identity of Baume LJ, Meyer J. 1966. Dental dysplasia related
domestication in northern China. Proceedings to malnutrition, with special reference to
of the National Academy of Sciences melanodontia and odontoclasia. Journal of
106:5523–5528. Dental Research 45:726–741.
Bartsiokas A, Day MH. 1993. Lead poisoning and Baynes RD, Bothwell TH. 1990. Iron deficiency.
dental caries in the Broken Hill hominid. Annual Review of Nutrition 10:133–148.
Journal of Human Evolution 24:243–249. Beals KL. 1972. Head form and climatic stress.
Bass SL, Saxon L, Daly RM, et al. 2002. The effect American Journal of Physical Anthropology
of mechanical loading on the size and shape of 37:85–92.
bone in pre-, peri-, and postpubertal girls: a Beals KL, Smith CL, Dodd SM. 1984. Brain size,
study in tennis players. Journal of Bone and cranial morphology, climate, and time
Mineral Research 17:2274–2280. machines. Current Anthropology 25:301–330.
Bass WM. 1964. The Variation in Physical Types Beck LA, editor. 1995. Regional Approaches to
of the Prehistoric Plains Indians. Plains Mortuary Analysis. New York, NY: Plenum
Anthropologist Memoir, No. 1. Press.
Bass WM. 2005. Human Osteology: A Laboratory Beck TJ, Ruff CB, Warden KE, Scott WW Jr., Rao
and Field Manual of the Human Skeleton, Fifth GU. 1990. Predicting femoral neck strength
Edition. Columbia, MO: Missouri from bone mineral data: a structural approach.
Archaeological Society. Investigative Radiology 25:6–18.
Basset A, Malgras J, Maleville J. 1994. Aspects des Becker MJ. 1993. Human skeletons from
tréponématoses endémiques non vénériennes Tarquinia: a preliminary analysis of the
en Afrique de l’ouest. In: Dutour O, Pálfi G, 1989 Cimitero site excavations with inferences
Bérato J, Brun JP, editors. L’Origine de la for the evaluation of Etruscan social classes.
Syphilis en Europe: Avant ou Après 1493? Studi Etruschi 58(3):211–248.
Paris, France: Éditions Errance; p. 55–56. Becker NSA, Verdu P, Froment A, et al. 2011.
Bassett EJ. 1981. Tetracycline staining in a Indirect evidence for the genetic determination
pre-antibiotic population from Sudanese of short stature in African Pygmies. American
Nubia. Bulletins et Mémoires de la Société‚ Journal of Physical Anthropology 145:390–401.
d’Anthropologie de Paris 8:321–322. Beecher RM, Corruccini RS. 1981. Effects of dietary
Bassett EJ. 1982. Osteological analysis of Carrier consistency on maxillary arch breadth in
Mills burials. In: Jefferies RW, Butler B, editors. macaques. Journal of Dental Research 60:68.
The Carrier Mills Archaeological Project: Beecher RM, Corruccini RS. 1983. Craniofacial
Human Adaptation in the Saline Valley, correlates of dietary consistency in a
Illinois, Volume II. Southern Illinois University nonhuman primate. Journal of Craniofacial
at Carbondale, Center for Archaeological Genetics and Developmental Biology 3:
Investigations, Research Paper, No. 33; 193–202.
p. 1027–1114. Begg PR. 1954. Stone age man’s dentition.
Bassett EJ, Keith MS, Armelagos GJ, Martin DL. American Journal of Orthodontics 40:298–312,
1980. Tetracycline-labeled human bone from 373–383, 462–475, 517–531.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 441

Behrend GA. 1978. The epidemiology of dental Paleopathology at the Origins of Agriculture.
caries and subsistence pattern change. Orland, FL: Academic Press; p. 531–558.
American Journal of Physical Anthropology Benfer RA, Edwards DS. 1991. The principal axis
48:380. method for measuring rate and amount of
Behrents RG. 1985. Growth in the Aging dental attrition: estimating juvenile or adult
Craniofacial Skeleton. University of Michigan tooth wear from unaged adult teeth. In: Kelley
Center for Human Growth and Development, MA, Larsen CS, editors. Advances in Dental
Craniofacial Growth Series Monograph, No. 17. Anthropology. New York, NY: Wiley-Liss;
Belcastro MG, Mariotti V, Facchini F, Dutour O. p. 325–340.
2005. Leprosy in a skeleton from the 7th Ben-Itzhak S, Smith P, Bloom RA. 1988.
century necropolis of Vicenne-Campochiaro Radiographic study of the humerus in
(Molise, Italy). International Journal of Neandertals and Homo sapiens sapiens.
Osteoarchaeology 15:431–448. American Journal of Physical Anthropology
Bellantoni NF, Sledzik PS, Poirier DA. 1997. 77:231–242.
Rescue, research, and reburial: Walton family Benjamin M, Toumi H, Ralphs JR, et al. 2006.
cemetery, Griswold, Connecticut. In: Poirier Where tendons and ligaments meet bone:
DA, Bellantoni NF, editors. In Remembrance: attachment sites (‘entheses’) in relation to
Archaeology and Death. Westport, CT: Bergin & exercise and/or mechanical load. Journal of
Garvey; p. 131–154. Anatomy 208:471–490.
Bello SM, Thomann A, Signoli M, Dutour O, Bennett KA. 1973. The Indians of Point of Pines,
Andrews P. 2006. Age and sex bias in the Arizona: A Comparative Study of their Physical
reconstruction of past population structures. Characteristics. Anthropological Papers of the
American Journal of Physical Anthropology University of Arizona, No. 23.
129:24–38. Bennett KA, Cheverud JM, Booth SN. 1981.
Bellwood P. 2001. Early agriculturalist population Deciduous tooth dimension in fetal rhesus
diasporas? Farming, languages, and genes. monkeys from mothers with induced diabetes.
Annual Review of Anthropology 30:181–207. American Journal of Physical Anthropology
Beltrán-Sánchez H, Crimmins EM, Finch CE. 2012. 55:411–417.
Early cohort mortality predicts the rate of Bennike P. 1985. Palaeopathology of Danish
aging in the cohort: a historical analysis. Skeletons: A Comparative Study of
Journal of Developmental Origins of Health and Demography, Disease and Injury. Copenhagen,
Disease 3:380–386. Denmark: Akademisk Forlag.
Bement LC. 1994. Hunter-Gatherer Mortuary Bennike P. 1991a. Epidemiological aspects of
Practices during the Central Texas Archaic. paleopathology in Denmark: past, present, and
Austin, TX: University of Texas Press. future studies. In: Ortner DJ, Aufderheide AC,
Ben-David Y, Hershkovitz I, Rubin D, Moscona D, editors. Human Paleopathology: Current
Ring B. 1992. Inbreeding effects on tooth size, Syntheses and Future Options. Washington,
eruption age and dental directional and DC: Smithsonian Institution Press; p. 140–144.
fluctuating asymmetry among South Sinai Bennike P. 1991b. Palaeopathological studies of
Bedouins. In: Smith P, Tchernov E, editors. the prehistoric and early historic population
Structure, Function and Evolution of Teeth. from Denmark. Veroff sbersee-Mus 9:
London, UK: Freund; p. 361–389. 129–146.
Bender MM, Baerreis DA, Steventon RL. 1981. Bennike P. 1999. Facts or myths? A re-evaluation
Further light on carbon isotopes and Hopewell of cases of diagnosed tuberculosis in Denmark.
agriculture. American Antiquity 46:346–353. In: Pálfi G, Dutour O, Deak J, Hutas I, editors.
Benfer RA. 1984. The challenges and rewards of Tuberculosis: Past and Present. Budapest/
sedentism: the preceramic village of Paloma, Szeged, Hungary: Golden Book Publishers and
Peru. In: Cohen MN, Armelagos GJ, editors. Tuberculosis Foundation; p. 511–518.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
442 References

Bennike P. 2002. Vilhelm Møller-Christensen: his human anterior teeth: evidence of handedness
work and legacy. In: Roberts CA, Lewis ME, in the middle and early Upper Pleistocene.
Manchester K, editors. The Past and Present of Journal of Human Evolution 17:403–412.
Leprosy: Archaeological, Historical, Bermúdez de Castro JM, Nicolas ME. 1995.
Palaeopathological and Clinical Approaches. Posterior dental size reduction in hominids: the
British Archaeological Reports, International Atapuerca evidence. American Journal of
Series, No. 1054; p. 135–144. Physical Anthropology 96:335–356.
Bennike P, Alexandersen V. 2007. Population Bermúdez de Castro JM, Pérez PJ. 1995. Enamel
plasticity in southern Scandinavia: from hypoplasia in the Middle Pleistocene hominids
oysters and fish to gruel and meat. In: Cohen from Atapuerca (Spain). American Journal of
MN, Crane-Kramer GMM, editors. Ancient Physical Anthropology 96:301–314.
Health: Skeletal Indicators of Agricultural and Berndt CH. 1978. In Aboriginal Australia. In:
Economic Intensification. Gainesville, FL: Montagu A, editor. Learning Non-Aggression:
University Press of Florida; p. 130–148. The Experience of Non-Literate Societies. New
Bennike P, Ebbesen K, Jorgensen LB. 1986. The York, NY: Oxford University Press; p. 144–160.
bog find from Sigersdal. Journal of Danish Berney S, Goldstein M, Bishko F. 1972. Clinical
Archaeology 5:85–115. and diagnostic features of tuberculous arthritis.
Benson EP, Boone EP, editors. 1984. Ritual Human American Journal of Medicine 53:36–42.
Sacrifice in Mesoamerica. Washington, DC: Berry AC, Berry RJ. 1967. Epigenetic variation in
Dumbarton Oaks. the human cranium. Journal of Anatomy
Bentley GR. 1991. A bioarchaeological 101:361–379.
reconstruction of the social and kinship Berry AC, Berry RJ. 1972. Origins and
systems at Early Bronze Age Bab edh-Dhra’, relationships of the ancient Egyptians, based
Jordan. In: Gregg SA, editor. Between Bands on a study of non-metrical variations in the
and States. Southern Illinois University at skull. Journal of Human Evolution 1:199–208.
Carbondale, Center for Archaeological Bhargava A. 1999. Modeling the effects of
Investigations, Occasional Paper, No. 9; nutritional and socioeconomic factors on the
p. 5–34. growth and morbidity of Kenya school
Bentley GR, Jasinska G, Goldberg T. 1993. Is the children. American Journal of Physical
fertility of agriculturalists higher than that of Anthropology 11:317–326.
nonagriculturalists? Current Anthropology Bianucci R, Rahalison L, Massa ER, et al. 2008.
34:778–785. A rapid diagnostic test detects plague in
Bentley RA. 2006. Strontium isotopes from the ancient human remains: an example of the
earth to the archaeological skeleton: a review. interaction between archeological and
Journal of Archaeological Method and Theory biological approaches (southeastern France,
13:135–187. 16th–18th century). American Journal of
Bentley RA, Pietrusewsky M, Douglas MT, Physical Anthropology 136:361–367.
Atkinson TC. 2005. Matrilocality during the Bianucci R, Rahalison L, Peluso A, et al. 2009.
prehistoric transition to agriculture in Plague immunodetection in remains of
Thailand? Antiquity 79:865–881. religious exhumed from burial sites in France.
Bentley RA, Price TD, Luning J, et al. 2002. Human Journal of Archaeological Science 36:
migration in early Neolithic Europe. Current 616–621.
Anthropology 43:799–804. Bielicki T, Welon Z. 1982. Growth data as
Berger TD, Trinkaus E. 1995. Patterns of trauma indicators of social inequalities: the case of
among the Neandertals. Journal of Poland. Yearbook of Physical Anthropology
Archaeological Science 22:841–852. 25:153–167.
Bermúdez de Castro JM, Bromage TG, Fernández Biggerstaff RH. 1970. Morphological variation for
Jalvo Y. 1988. Buccal striations on fossil the permanent mandibular first molars in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 443

human monozygotic and dizygotic twins. Blakely RL. 1977. Sociocultural implications of
Archives of Oral Biology 15:721–733. demographic data from Etowah, Georgia. In:
Biknevicius AR, Ruff CB. 1992. Use of biplanar Blakely RL, editor. Biocultural Adaptation in
radiographs for estimating cross-sectional Prehistoric America. Athens, GA: University of
geometric properties of mandibles. Anatomical Georgia Press; p. 45–66.
Record 232:157–163. Blakely RL. 1980. Sociocultural implications of
Billman BR. 2008. An outbreak of violence and pathology between the village area and Mound
raiding in the central Mesa Verde region in the C skeletal remains from Etowah, Georgia. In:
12th century AD. In: Nichols DL, Crown PL, Willey P, Smith PH, editors. The Skeletal
editors. Social Violence in the Prehispanic Biology of Aboriginal Populations in the
American Southwest. Tucson, AZ: University Southeastern United States. Tennessee
of Arizona Press; p. 41–69. Anthropological Association, Miscellaneous
Billman BR, Lambert PM, Leonard BL. 2000. Paper, No. 5; p. 28–38.
Cannibalism, warfare, and drought in the Mesa Blakely RL. 1988. The life cycle and social
Verde region during the twelfth century A.D. organization. In: Blakely RL, editor. The King
American Antiquity 65:145–178. Site: Continuity and Contact in Sixteenth-
Bird HA, Eastmond CJ, Hudson A, Wright V. 1980. Century Georgia. Athens, GA: University of
Is generalized joint laxity a factor in Georgia Press; p. 17–34.
spondylolisthesis? Scandinavian Journal of Blakely RL. 1989. Bone strontium in pregnant and
Rheumatology 9:203–205. lactating females from archaeological samples.
Birkby WH. 1982. Biosocial interpretations from American Journal of Physical Anthropology
cranial nonmetric traits of Grasshopper Pueblo 80:173–186.
skeletal remains. In: Longacre WA, Holbrook Blakely RL. 1995. Social organization at Etowah:
SJ, Graves MW, editors. Multidisciplinary a reconsideration of paleodemographic and
Research at Grasshopper Pueblo, Arizona. paleonutritional evidence. Southeastern
Anthropological Papers of the University of Archaeology 14:46–59.
Arizona, No. 40; p. 36–41. Blakely RL, Beck LA. 1981. Trace elements,
Black TK. 1980. An exception to the apparent nutritional status, and social stratification at
relationships between stress and fluctuating Etowah, Georgia. Annals of the New York
dental asymmetry. Journal of Dental Research Academy of Sciences 376:417–431.
59:1168–1169. Blakely RL, Beck L. 1984. Tooth-tool use versus
Blackburn A. 2011. Bilateral asymmetry of the dental mutilation: a case study from the
humerus during growth and development. prehistoric Southeast. Midcontinental Journal
American Journal of Physical Anthropology of Archaeology 9:269–277.
145:639–646. Blakely RL, Mathews DS. 1990. Bioarchaeological
Blair EH. 2013. The Guale landscape of Santa evidence for a Spanish – Native American
Catalina de Guale: 30 years of geophysics at a conflict in the sixteenth-century Southeast.
Spanish colonial mission. In: Thompson VD, American Antiquity 55:718–744.
Thomas DH, editors. Life among the Tides: Blakeslee DJ. 1994. The archaeological context of
Recent Archaeology on the Georgia Bight. human skeletons in the Northern and Central
Anthropological Papers of the American Plains. In: Owsley DW, Jantz RL, editors.
Museum of Natural History, No. 98; Skeletal Biology in the Great Plains: Migration,
p. 375–393. Warfare, Health, and Subsistence. Washington,
Blakely RL. 1971. Comparison of the mortality DC: Smithsonian Institution Press; p. 9–32.
profiles of Archaic, Middle Woodland, and Blakey ML. 2001. Bioarchaeology of the African
Middle Mississippian skeletal populations. diaspora in the Americas: its origins and
American Journal of Physical Anthropology scope. Annual Review of Anthropology
34:43–54. 30:387–422.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
444 References

Blakey ML. 2010. African Burial Ground Project: Block JA, Shakoor N. 2010. Lower limb
paradigm for cooperation? Museum osteoarthritis: biomechanical alterations and
International 62:61–68. implications for therapy. Current Opinion in
Blakey ML, Armelagos GJ. 1985. Deciduous dental Rheumatology 22:544–550.
enamel defects in prehistoric Americans from Blom DE, Buikstra JE, Keng L, et al. 2005. Anemia
Dickson Mounds: prenatal and postnatal stress. and childhood mortality: latitudinal patterning
American Journal of Physical Anthropology along the coast of pre-Columbian Peru.
66:371–380. American Journal of Physical Anthropology
Blakey M, Leslie TE, Reidy JP. 1994. Frequency 127:152–169.
and chronological distribution of dental Blom DE, Janusek JW, Buikstra JE. 2003.
enamel hypoplasia in enslaved African A reevaluation of the human remains from
Americans: a test of the weaning hypothesis. Tiwanaku. In: Kolata A, editor. Tiwanaku and
American Journal of Physical Anthropology Its Hinterland: Archaeological and
95:371–383. Paleoecological Investigations of an Andean
Blanco RA, Acheson RM, Canosa C, Salomon JB. Civilization, Volume 2. Urban and Rural
1974. Height, weight, and lines of arrested Archaeology. Washington, DC: Smithsonian
growth in young Guatemalan children. Institution Press; p. 435–448.
American Journal of Physical Anthropology Blondiaux J, Durr J, Khouchaf L, Eisenberg LE.
40:39–48. 2002. Microscopic study and x-ray analysis of
Blanton RE. 1978. Monte Albán: Settlement two 5th century cases of leprosy:
Patterns at the Ancient Zapotec Capital. New palaeoepidemiological and clinical approaches.
York, NY: Academic Press. In: Robert CA, Lewis ME, Manchester K,
Blanton RE. 1981. The rise of cities. In: Sabloff JA, editors. The Past and Present of Leprosy:
editor. Supplement to the Handbook of Middle Archaeological, Historical, Palaeopathological
American Indians. Austin, TX: University of and Clinical Approaches. British
Texas Press; p. 392–400. Archaeological Reports, International Series,
Blatt, SH. 2013. From the Mouth of Babes: No. 1054; p. 105–110.
Reconstructing Dental Age and Growth of Blondiaux J, Duvette J-F, Eisenberg LE. 1994.
Prehistoric Native American Juveniles Using Microradiographs of leprosy from an
Incremental Microstructures of Enamel. PhD osteoarchaeological context. International
Dissertation, Ohio State University, Columbus, OH. Journal of Osteoarchaeology 4:13–20.
Blau S. 2007. Skeletal and dental health and Blondiaux J, Hedain P, Chastanet M, et al. 1999.
subsistence change in the United Arab Epidemiology of tuberculosis: a 4th to 12th c. AD
Emirates. In: Cohen MN, Crane-Kramer GMM, picture in a 2498-skeleton series from northern
editors. Ancient Health: Skeletal Indicators of France. In: Pálfi G, Dutour O, Deak J, Hutas I,
Agricultural and Economic Intensification. editors. Tuberculosis: Past and Present. Budapest/
Gainesville, FL: University Press of Florida; Szeged, Hungary: Golden Book Publishers and
p. 190–206. Tuberculosis Foundation; p. 521–530.
Blau S, Yagodin V. 2005. Osteoarchaeological Bloom AI, Bloom RA, Kahila G, Eisenberg E, Smith
evidence for leprosy from western central Asia. P. 1995. Amputation of the hand in the 3600-
American Journal of Physical Anthropology year-old skeletal remains of an adult male: the
126:150–158. first case reported from Israel. International
Blinkhorn AS, Davies RM. 1996. Caries Journal of Osteoarchaeology 5:188–191.
prevention: a continued need worldwide. Boas F. 1902. Rudolf Virchow’s anthropological
International Dental Journal 46:119–125. work. Science 16:441–445.
Blitz JH. 1988. Adoption of the bow in prehistoric Boas F. 1912. Changes in bodily form of
North America. North American Archaeologist descendants of immigrants. American
9:123–145. Anthropologist 14:530–563.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 445

Boas F. 1916. New evidence in regard to the Demography in the Americas. Washington, DC:
instability of human types. Proceedings of the Smithsonian Institution Press; p. 155–163.
National Academy of Sciences 2:713–718. Bogin B, Keep R. 1999. Eight thousand years of
Bocquentin F. 2007. A Final Natufian population: economic and political history in Latin
health and burial status at Eynan-Mallaha. America revealed by anthropometry. Annals of
In: Faerman M, Horwitz LK, Kahana T, Human Biology 26:333–351.
Zilberman U, editors. Faces from the Past: Bogin BA. 1999. Patterns of Human Growth,
Diachronic Patterns in the Biology of Human Second Edition. Cambridge, UK: Cambridge
Populations from the Eastern Mediterranean: University Press.
Papers in Honour of Patricia Smith. British Bogin BA, Loucky J. 1997. Plasticity, political
Archaeological Reports, International Series, economy, and physical growth status of
No. 1603; p. 67–81. Guatemala Maya children living in the United
Bocquet-Appel J-P. 2002. Paleoanthropological States. American Journal of Physical
traces of Neolithic demographic transition. Anthropology 102:17–32.
Current Anthropology 43:637–650. Bogin BA, MacVean RB. 1978. Growth in height
Bocquet-Appel J-P, editor. 2008. Recent Advances and weight of urban Guatemalan primary
in Paleodemography: Data, Techniques, school children of low and high socioeconomic
Patterns. Dordrecht, the Netherlands: Springer. class. Human Biology 50:477–487.
Bocquet-Appel J-P. 2011. When the world’s Bogin BA, MacVean RB. 1981. Nutritional and
population took off: the springboard of the biological determinants of body fat patterning
Neolithic Demographic Transition. Science in urban Guatemalan children. Human Biology
333:560–561. 53:259–268.
Bocquet-Appel J-P, Bar-Yosef O, editors. 2008. Bogin BA, MacVean RB. 1983. The relationship of
The Neolithic Demographic Transition and its socioeconomic status and sex to body size,
Consequences. New York, NY: Springer. skeletal maturation, and cognitive status of
Bocquet-Appel J-P, Masset C. 1982. Farewell to Guatemala City schoolchildren. Child
paleodemography. Journal of Human Evolution Development 54:115–128.
11:321–333. Boklage CE. 1987. Developmental differences
Bocquet-Appel J-P, Naji S. 2006. Testing the between singletons and twins in distributions
hypothesis of a worldwide Neolithic of dental diameter asymmetries. American
demographic transition: corroboration from Journal of Physical Anthropology 74:319–331.
American cemeteries. Current Anthropology Boldsen JL. 1995. The place of plasticity in the study
47:341–365. of the secular trend for male stature: an analysis
Bocquet-Appel J-P, Naji S, Bandy M. 2008. of Danish biological population history. In:
Demographic and health changes during the Mascie-Taylor CGN, Bogin B, editors. Human
transition to agriculture in North America. In: Variability and Plasticity. Cambridge, UK:
Bocquet-Appel J-P, editor. Recent Advances in Cambridge University Press; p. 75–90.
Paleodemography: Data, Techniques, Patterns. Boldsen JL. 2005. Leprosy and mortality in the
Dordrecht, the Netherlands: Springer; Medieval Danish village of Tirup. American
p. 277–292. Journal of Physical Anthropology 126:159–168.
Bogaard A, Fraser R, Heaton THE, et al. 2013. Crop Boldsen JL. 2007. Early childhood stress and adult
manuring and intensive land management age mortality: A study of dental enamel
by Europe’s first farmers. Proceedings of the hypoplasia in the medieval village of Tirup.
National Academy of Sciences 110: American Journal of Physical Anthropology
12589–12594. 132:59–66.
Bogdan G, Weaver DS. 1992. Pre-Columbian Boldsen JL. 2008. Leprosy in the early Medieval
treponematosis in coastal North Carolina. In: Lauchheim community. American Journal of
Verano JW, Ubelaker DH, editors. Disease and Physical Anthropology 135:301–310.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
446 References

Boldsen JL, Milner GR. 2012. An epidemiological Borgognini Tarli S, Silvana M, Repetto E. 1985.
approach to paleopathology. In: Grauer A, Diet, dental features and oral pathology in the
editor. A Companion to Paleopathology. Mesolithic samples from Uzzo and Molara
Chichester, UK: Wiley-Blackwell; p. 114–132. Caves (Sicily). In: Malon C, Stoddart S, editors.
Boldsen JL, Milner GR, Konigsberg LW, Wood JW. Papers in Italian Archaeology IV (II):
2002. Transition analysis: a new method for Prehistory. British Archaeological Reports,
estimating age from skeletons. In: Hoppa RD, International Series, No. 244; p. 87–100.
Vaupel JW, editors. Paleodemography: Age Borgognini Tarli SM, Repetto E. 1986. Skeletal
Distributions from Skeletal Samples. Cambridge, indicators of subsistence patterns and activity
UK: Cambridge University Press; p. 73–106. regime in the Mesolithic sample from Grotta
Boldsen JL, Mollerup L. 2006. Outside St. Jorgen: dell’Uzzo (Trapani, Sicily): a case study. Human
leprosy in the Medieval Danish city of Odense. Evolution 1:1–21.
American Journal of Physical Anthropology Boros-Major A, Bono A, Lovasz G, et al. 2011.
130:344–351. New perspectives in biomolecular
Bolnick DA, Smith DG. 2007. Migration and social paleopathology of ancient tuberculosis: a
structure among the Hopewell: evidence from proteomic approach. Journal of Archaeological
ancient DNA. American Antiquity 72:627–644. Science 38:197–201.
Bolton HE. 1927. Fray Juan Crespí, Missionary Bos KI, Harkins KM, Herbig A, et al. 2014.
Explorer on the Pacific Coast, 1769–1774. Pre-Columbian mycobacterial genomes reveal
Berkeley, CA: University of California Press. seals as a source of New World tuberculosis.
Bonassie P. 1989. Consommation d’aliments Nature 514:494–497.
immondes et cannibalisme de–survie dans Bos KI, Schuenemann VJ, Golding GB, et al. 2011.
l’occident du haut moyen âge. Annales ESC A draft genome of Yersinia pestis from victims
5:1035–1056. of the Black Death. Nature 478:506–510.
Bonatto SL, Salzano FM. 1997. A single and early Botella MC, Alemán I, Jiménez SA. 2000. Los
migration for the peopling of the Americas Huesos Humanos: Manipulatión y Altraciones.
supported by mitochondrial DNA sequence Barcelona, Spain: Edicións Bellaterra.
data. Proceedings of the National Academy of Boulestin B, Zeeb-Lanz A, Jeunesse C, et al. 2009.
Sciences 94:1866–1871. Mass cannibalism in the Linear Pottery culture
Bonfiglioli B, Mariotti V, Facchini F, Belcastro at Herxheim (Palatinate, Germany). Antiquity
MG. 2004. Masticatory and non-masticatory 83:968–982.
dental modifications in the Epipalaeolithic Boulle E-L. 2001. Evolution of two human skeletal
necropolis of Taforalt (Morocco). International markers of the squatting position: a diachronic
Journal of Osteoarchaeology 14:448–456. study from antiquity to the modern age.
Bonsall C, Cook GT, Hedges REM, et al. 2004. American Journal of Physical Anthropology
Radiocarbon and stable isotope evidence of 115:50–56.
dietary change from the Mesolithic to the Bourdieu P. 1977. Outline of a Theory of Practice.
Middle Ages in the Iron Gates: new results from Cambridge, UK: Cambridge University Press.
Lepenski Vir. Radiocarbon 46:293–300. Bourque BJ, Krueger HW. 1994. Dietary
Boocock P, Roberts C, Manchester K. 1995. reconstruction from human bone isotopes for
Maxillary sinusitis in Medieval Chichester, five coastal New England populations. In:
England. American Journal of Physical Sobolik KD, editor. Paleonutrition: The Diet
Anthropology 98:483–495. and Health of Prehistoric Americans. Southern
Borgognini Tarli S, Canci A, Piperno M, Repetto E. Illinois University at Carbondale, Center for
1993. Dati archeologici e antropologici sulle Archaeological Investigations, Occasional
sepolture mesolithiche della Grotta dell’Uzzo Paper, No. 22; p. 195–209.
(Trapini). Bullettino di Paletnologia Italiana Boutton TW, Lynott MJ, Bumsted MP. 1991.
84:85–179. Stable carbon isotopes and the study of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 447

prehistoric human diet. Critical Reviews in from a Selected Southeastern Indian Skeletal
Food Science and Nutrition 30:373–385. Temporal Series. PhD Dissertation, University
Bouvier M, Hylander WL. 1981. Effect of bone of Tennessee, Knoxville, TN.
strain on cortical bone structure in macaques Brabant H. 1967. Palaeostomatology. In:
(Macaca mulatta). Journal of Morphology Brothwell DR, Sandison AT, editors. Diseases in
167:1–12. Antiquity. Springfield, IL: Charles C. Thomas;
Bouvier M, Hylander WL. 1982. The effect of p. 538–550.
dietary consistency on gross and histologic Brace CL. 1995. Biocultural interaction and the
morphology in the craniofacial region of mechanism of mosaic evolution in the
young rats. American Journal of Anatomy emergence of “modern” morphology. American
170:117–126. Anthropologist 97:711–721.
Bouvier M, Hylander WL. 1984. The effect of Brace CL, Hinton RJ. 1981. Oceanic tooth-size
dietary consistency on morphology of the variation as a reflection of biological and
mandibular condylar cartilage in young cultural mixing. Current Anthropology
macaques (Macaca mulatta). In: Dixon AD, 22:549–569.
Sarnat BG, editors. Factors and Mechanisms Brace CL, Mahler PE. 1971. Post-Pleistocene
Influencing Bone Growth. New York, NY: Alan changes in the human dentition. American
R. Liss; p. 569–579. Journal of Physical Anthropology 34:191–204.
Bouvier M, Zimny ML. 1987. Effects of mechanical Brace CL, Nagai M. 1982. Japanese tooth size: past
loads on surface morphology of the condylar and present. American Journal of Physical
cartilage of the mandible in rats. Acta Anthropology 59:399–411.
Anatomica 129:293–300. Brace CL, Rosenberg KR, Hunt KD. 1987. Gradual
Bouwman AS, Brown TA. 2005. The limits of change in human tooth size in the late
biomolecular palaeopathology: ancient DNA Pleistocene and post Pleistocene. Evolution
cannot be used to study venereal syphilis. 41:705–720.
Journal of Archaeological Science 32:702–713. Brace CL, Smith SL, Hunt KD. 1991. What big teeth
Bovee DL, Owsley DW. 1994. Evidence of warfare you had grandma! Human tooth size, past and
at the Heerwald site. In: Owsley DW, Jantz RL, present. In: Kelley MA, Larsen CS, editors.
editors. Skeletal Biology in the Great Plains: Advances in Dental Anthropology. New York,
Migration, Warfare, Health, and Subsistence. NY: Wiley-Liss; p. 33–57.
Washington, DC: Smithsonian Institution Braidwood RJ. 1967. Prehistoric Men, Seventh
Press; p. 355–362. Edition. Glenview, IL: Scott, Foresman.
Bowen WH. 1994. Food components and caries. Brandt G, Haak W, Adler CJ, et al. 2013. Ancient
Advances in Dental Research 8:215–220. DNA reveals key stages in the formation of
Boyd DC. 1986. A comparison of Mouse Creek central European mitochondrial genetic
phase to Dallas and Middle Cumberland culture diversity. Science 342:257–261.
skeletal remains. In: Levy JE, editor. Skeletal Braun M, Cook DC, Pfeiffer S. 1998. DNA from
Analysis in Southeastern Archaeology. North Mycobacterium tuberculosis complex identified
Carolina Archaeological Council Publication, in North American, pre-Columbian human
No. 24; p. 103–126. skeletal remains. Journal of Archaeological
Boyd DC, Boyd CC Jr. 1989. A comparison of Science 25:271–277.
Tennessee Archaic and Mississippian maximum Brickley M. 2006. Rib fractures in the
femoral lengths and midshaft diameters: archaeological record: a useful source of
subsistence change and postcranial variability. sociocultural information? International
Southeastern Archaeology 8:107–116. Journal of Osteoarchaeology 16:61–75.
Boyd DCM. 1988. A Functional Model for Brickley M, Ives R. 2006. Skeletal manifestations
Masticatory-Related Mandibular, Dental, and of infantile scurvy. American Journal of
Craniofacial Microevolutionary Change Derived Physical Anthropology 129:163–172.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
448 References

Brickley M, Ives R. 2008. The Bioarchaeology of Bridges PS. 1994. Vertebral arthritis and physical
Metabolic Bone Disease. Oxford, UK: Elsevier activities in the prehistoric southeastern United
Academic Press. States. American Journal of Physical
Brickley M, Mays S, Ives R. 2005. Skeletal Anthropology 93:83–93.
manifestations of vitamin D deficiency Bridges PS. 1995. Skeletal biology and behavior in
osteomalacia in documented skeletal ancient humans. Evolutionary Anthropology
collections. International Journal of 4:112–120.
Osteoarchaeology 15:389–403. Bridges PS. 1996. Warfare and mortality at
Brickley M, Mays S, Ives R. 2007. An investigation Koger’s Island, Alabama. International Journal
of skeletal indicators of vitamin D deficiency in of Osteoarchaeology 6:66–75.
adults: effective markers for interpreting past Bridges PS, Blitz JH, Solano MC. 2000. Changes in
living conditions and pollution levels in 18th long bone diaphyseal strength with
and 19th century Birmingham, England. horticultural intensification in west-central
American Journal of Physical Anthropology Illinois. American Journal of Physical
132:67–79. Anthropology 112:217–238.
Bridges PS. 1989a. Spondylolysis and its Bridges PS, Jacobi KP, Powell ML. 2000. Warfare-
relationship to degenerative joint disease in the related trauma in the late prehistory of Alabama.
prehistoric southeastern United States. In: Lambert PM, editor. Bioarchaeological
American Journal of Physical Anthropology Studies in the Age of Agriculture: A View from
79:321–329. the Southeast. Tuscaloosa, AL: University of
Bridges PS. 1989b. Changes in activities with the Alabama Press; p. 35–62.
shift to agriculture in the southeastern United Brinch O, Møller-Christensen V. 1949. On
States. Current Anthropology 30:385–394. comparative investigations into the occurrence
Bridges PS. 1990. Osteological correlates of of dental caries in archaeological skulls.
weapon use. In: Buikstra JE, editor. A Life in Odontologisk Tidskrift 57:357–373.
Science: Papers in Honor of J. Lawrence Angel. Brinker RA. 1985. Pelvic index in Illinois
Center for American Archeology, Scientific Hopewell. American Journal of Physical
Papers, No. 6; p. 87–98. Anthropology 66:150.
Bridges PS. 1991a. Degenerative joint disease Brock SL. 1985. Biomechanical Adaptation of the
in hunter-gatherers and agriculturalists from Lower Limb Bones through Time in the
the southeastern United States. American Prehistoric Southwest. PhD Dissertation,
Journal of Physical Anthropology 85: University of New Mexico, Albuquerque, NM.
379–391. Brock SL, Ruff CB. 1988. Diachronic patterns of
Bridges PS. 1991b. Skeletal evidence of changes in change in structural properties of the femur in
subsistence activities between the Archaic and the prehistoric American Southwest. American
Mississippian time periods in northwestern Journal of Physical Anthropology 75:113–127.
Alabama. In: Powell ML, Bridges PS, Mires Brooks ST. 1955. Skeletal age at death: the
AMW, editors. What Mean These Bones? reliability of cranial and pubic age indicators.
Studies in Southeastern Bioarchaeology. American Journal of Physical Anthropology
Tuscaloosa, AL: University of Alabama Press; 13:567–597.
p. 89–101. Brosch R, Gordon SV, Marmiesse M, et al. 2002.
Bridges PS. 1992. Prehistoric arthritis in the A new evolutionary scenario for the
Americas. Annual Review of Anthropology Mycobacterium tuberculosis complex.
21:67–91. Proceedings of the National Academy of
Bridges PS. 1993. The effect of variation in Sciences 99:3684–3689.
methodology on the outcome of osteoarthritic Brothwell DR. 1959. Teeth in earlier human
studies. International Journal of populations. Proceedings of the Nutrition
Osteoarchaeology 3:289–295. Society 18:59–65.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 449

Brothwell DR. 1961. The palaeopathology of early Brown JA. 1971. Approaches to the Social
British man: an essay on the problems of Dimensions of Mortuary Practices. Memoirs of
diagnosis and analysis. Journal of the Royal the Society for American Archaeology, No. 25.
Anthropological Institute 91:318–344. Brown KA, Brown TA. 2013. Biomolecular
Brothwell D. 1971. Forensic aspects of the archaeology. Annual Review of Anthropology
so-called Neolithic skeleton Q1 from Maiden 42:159–174.
Castle, Dorset. World Archaeology 3:233–41. Brown PJ, Armelagos GJ, Maes KC. 2011. Humans
Brothwell DR. 1981. Digging up Bones: The in a world of microbes: the anthropology of
Excavation, Treatment and Study of Human infectious disease. In: Singer M, Erickson PI,
Skeletal Remains, Third Edition. Ithaca, NY: editors. A Companion to Medical
Cornell University Press. Anthropology. Chichester, UK: Wiley-
Brothwell DR. 1987. Decay and disorder in the Blackwell; p. 253–269.
York Jewbury skeletons. In: Boddington A, Brown T. 1992. Developmental, morphological
Garland AN, Janaway RC, editors. Death, and functional aspects of occlusion in
Decay and Reconstruction: Approaches to Australian Aboriginals. In: Lukacs JR, editor.
Archaeological and Forensic Science. Culture, Ecology and Dental Anthropology.
Manchester, UK: Manchester University Press; Journal of Human Ecology 2 (Special
p. 22–26. Issue):73–85.
Brothwell DR. 2005. North American Brown T, Brown K. 2011. Biomolecular
treponematosis against the bigger world Archaeology: An Introduction. Chichester, UK:
picture. In: Powell ML, Cook DC, editors. The Wiley-Blackwell.
Myth of Syphilis: The Natural History of Brown T, Molnar S. 1990. Interproximal grooving
Treponematosis in North America. Gainesville, and task activity in Australia. American
FL: University Press of Florida; p. 480–496. Journal of Physical Anthropology 81:
Brothwell DR, Powers R, Hirst S. 2000. The 545–553.
pathology. In: Rahtz P, Hirst S, Wrights SM, Bruwelheide KS, Owsley DW, Jantz RL. 2010.
editors. Cannington Cemetery: Excavations Burials from the fourth mission church at
1962–3 of Prehistoric, Roman and Later Pecos. In: Morgan ME, editor. Pecos Pueblo
Features at Cannington Park Quarry, Near Revisited: The Biological and Social Context.
Bridgewater, Somerset. London, UK: Society for Cambridge, MA: Papers of the Peabody
the Promotion of Roman Studies; p. 195–256. Museum, Volume 85; p. 129–160.
Brower AC. 1977. Cortical defect of the humerus at Buckberry J. 2008. Off with their heads: the
the insertion of the pectoralis major. American Anglo-Saxon execution cemetery at
Journal of Roentgenology 128:677–678. Walkington Wold, East Yorkshire. In: Murphy
Browman DL. 1978. Towards the development of EM, editor. Deviant Burial in the
the Tihuanaco (Tiwanaku) state. In: Browman Archaeological Record. Oxford, UK: Oxbow
DL, editor. Advances in Andean Archaeology. Books; p. 148–168.
The Hague, the Netherlands: Mouton; p. 327– Buckland PC, Amorosi T, Barlow LK, et al. 1996.
349. Bioarchaeological and climatological evidence
Brown AB. 1973. Bone Strontium Content as a for the fate of Norse farmers in Medieval
Dietary Indicator in Human Skeletal Greenland. Antiquity 70:88–96.
Populations. PhD Dissertation, University of Bufkin WJ. 1971. The avulsive cortical
Michigan, Ann Arbor, MI. irregularity. American Journal of
Brown AB. 1988. Diet and nutritional stress. In: Roentgenology 112:487–492.
Blakely RL, editor. The King Site: Continuity Buhlin K, Gustafsson A, Pockley G, Frostegard J,
and Contact in Sixteenth-Century Georgia. Klinge B. 2003. Risk factors for cariovascular
Athens, GA: University of Georgia Press; disease in patients with periodontitis. European
p. 73–86. Heart Journal 24:2099–2107.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
450 References

Buhr AJ, Cooke AM. 1959. Fracture patterns. The University of Calgary, Proceedings of the 19th
Lancet 1:531–536. Annual Chacmool Conference; p. 243–259.
Buikstra JE. 1975. Cultural and biological Buikstra JE, Baker BJ, Cook DC. 1993. What
variability: a comparison of models. American diseases plagued ancient Egyptians? A century
Journal of Physical Anthropology 42:293. of controversy considered. In: Davis WV,
Buikstra JE. 1976a. Hopewell in the Lower Illinois Walker R, editors. Biological Anthropology and
Valley: A Regional Study of Human Biological the Study of Ancient Egypt. London, UK: British
Variability and Prehistoric Mortuary Behavior. Museum Press; p. 24–53.
Northwestern Archaeological Program Buikstra JE, Beck LA, editors. 2006.
Scientific Papers, No. 2. Bioarchaeology: The Contextual Analysis of
Buikstra JE. 1976b. The Caribou Eskimo: general Human Remains. Amsterdam, the Netherlands:
and specific disease. American Journal of Elsevier.
Physical Anthropology 45:351–368. Buikstra JE, Bullington J, Charles DK, et al. 1987.
Buikstra JE. 1977a. Biocultural dimensions of Diet, demography, and the development of
archeological study: a regional perspective. In: horticulture. In: Keegan WF, editor. Emergent
Blakely RL, editor. Biocultural Adaptation in Horticultural Economies of the Eastern
Prehistoric America. Athens, GA: University of Woodlands. Southern Illinois University at
Georgia Press; p. 67–84. Carbondale, Center for Archaeological
Buikstra JE. 1977b. Differential diagnosis: an Investigations Occasional Paper, No.7; p. 67–85.
epidemiological model. Yearbook of Physical Buikstra JE, Cook DC. 1978. Pre-Columbian
Anthropology 20:316–328. tuberculosis: an epidemiological approach.
Buikstra JE. 1980. Epigenetic distance: a study of Medical College of Virginia Quarterly 14:
biological variability in the lower Illinois River 32–44.
valley. In: Browman DL, editor. Early Native Buikstra JE, Cook DC. 1980. Palaeopathology: an
Americans: Prehistoric Demography, Economy American account. Annual Review of
and Technology. The Hague, the Netherlands: Anthropology 9:433–470.
Mouton; p. 271–299. Buikstra JE, Cook DC. 1981. Pre-Columbian
Buikstra JE. 1992. Diet and disease in late tuberculosis in west-central Illinois: prehistoric
prehistory. In: Verano JW, Ubelaker DH, disease in biocultural perspective. In: Buikstra
editors. Disease and Demography in the JE, editor. Prehistoric Tuberculosis in the
Americas. Washington, DC: Smithsonian Americas. Center for American Archeology,
Institution Press; p. 87–101. Scientific Papers, No. 5; p. 115–139.
Buikstra JE. 2006. Preface. In: Buikstra JE, Beck Buikstra JE, Frankenberg SR, Konigsberg LW.
LA, editors. Bioarchaeology: The Contextual 1990. Skeletal biological distance studies in
Analysis of Human Remains. Amsterdam, the American physical anthropology: recent
Netherlands: Academic Elsevier; p. xvii–xx. trends. American Journal of Physical
Buikstra JE. 2010. Paleopathology: a Anthropology 82:1–7.
contemporary perspective. In: Larsen CS, Buikstra JE, Frankenburg S, Lambert JB, Xue L.
editor. A Companion to Biological 1989. Multiple elements: multiple
Anthropology. Chichester, UK: Wiley- expectations. In: Price TD, editor. The
Blackwell; p. 395–411. Chemistry of Prehistoric Human Bone.
Buikstra JE, Autry W, Breitburg E, Eisenberg L, Cambridge, UK: Cambridge University Press;
van der Merwe N. 1988. Diet and health in the p. 155–210.
Nashville Basin: human adaptation and maize Buikstra JE, Konigsberg LW. 1985.
agriculture in Middle Tennessee. In: Kennedy Paleodemography: critiques and controversies.
BV, LeMoine GM, editors. Diet and American Anthropologist 87:316–333.
Subsistence: Current Archaeological Buikstra JE, Konigsberg LW, Bullington J. 1986.
Perspectives. Archaeological Association of the Fertility and the development of agriculture in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 451

the prehistoric Midwest. American Antiquity Population. Los Alamos, NM: Los Alamos
51:528–546. National Laboratory.
Buikstra JE, Milner GR. 1991. Isotopic and Burbank VK. 1994. Fighting Women: Anger and
archaeological interpretations of diet in the Aggression in Aboriginal Australia. Berkeley,
central Mississippi valley. Journal of CA: University of California Press.
Archaeological Science 18:319–329. Buretić-Tomljanović A, Ostojić S, Kapović M.
Buikstra JE, Ubelaker D, editors. 1994. Standards 2006. Secular change of craniofacial measures
for Data Collection from Human Skeletal in Croatian younger adults. American Journal
Remains. Arkansas Archaeological Survey of Human Biology 18:668–675.
Research Series, No. 44. Burger J, Thomas MG. 2011. The palaeopopulation
Buikstra JE, Williams S. 1991. Tuberculosis in the genetics of humans, cattle and dairying in
Americas: current perspectives. In: Ortner DJ, Neolithic Europe. In: Pinhasi R, Stock JT,
Aufderheide AC, editors. Human editors. Human Bioarchaeology of the
Paleopathology: Current Syntheses and Future Transition to Agriculture. Chichester, UK:
Options. Washington, DC: Smithsonian Wiley-Blackwell; p. 371–384.
Institution Press; p. 161–172. Burger RL, Lee-Thorp JA, van der Merwe NJ. 2003.
Bulbeck D, Lauer A. 2006. Human variation and Rite and crop in the Inca state revisited: an
evolution in Holocene peninsular Malaysia. In: isotopic perspective from Machu Picchu and
Oxenham M, Tayles N, editors. Bioarchaeology beyond. In: Burger RL, Salazar LC, editors. The
of Southeast Asia. Cambridge, UK: Cambridge 1912 Yale Peruvian Scientific Expedition
University Press; p. 133–171. Collections from Machu Picchu Human and
Bulger EM, Arneson MA, Mock CN, Jurkovich GJ. Animal Remains. Yale University Publications
2000. Rib fractures in the elderly. Journal of in Anthropology, No. 85; p. 119–137.
Trauma: Injury, Infection and Critical Care Burne R. 1998. Oral streptococci: products of their
48:1040–1046. environment. Journal of Dental Research
Bullen AK. 1972. Paleoepidemiology and 77:445–452.
distribution of prehistoric treponemiasis Burns PE. 1979. Log-linear analysis of dental
(syphilis) in Florida. Florida Anthropologist caries occurrence in four skeletal series.
25:133–174. American Journal of Physical Anthropology
Bullen AK. 1973. Some human skeletal remains 51:637–648.
from Amelia Island, Florida. In: Hemmings ET, Burns PE. 1982. A Study of Sexual Dimorphism in
Deagan KA, editors. Excavations on Amelia the Dental Pathology of Ancient Peoples. PhD
Island in Northeast Florida. Contributions of Dissertation, Arizona State University,
the Florida State Museum: Anthropology and Tempe, AZ.
History, No.18; p. 72–87. Burr DB. 1980. The relationship among physical,
Bullington J. 1991. Deciduous dental microwear of geometrical and mechanical properties of bone,
prehistoric juveniles from the lower Illinois with a note on the properties of nonhuman
River valley. American Journal of Physical primate bone. Yearbook of Physical
Anthropology 84:59–73. Anthropology 23:109–146.
Bullion SK. 1986. Information from teeth on the Burr DB, Allen MR, editors. 2013. Basic and
growth and developmental history of Applied Bone Biology. London, UK: Academic
individuals. In: Cruwys E, Foley RA, editors. Press.
Teeth and Anthropology. British Archaeological Burr DB, Forwood MR, Fyhrie DP, et al. 1997. Bone
Reports, International Series, No. 291; microdamage and skeletal fragility in
p. 133–136. osteoporotic and stress fractures. Journal of
Bumsted MP. 1984. Human Variation: δ13C in Bone and Mineral Research 12:6–15.
Adult Bone Collagen and the Relation to Diet in Burr DB, Ruff CB, Thompson DD. 1990. Patterns of
an Isochronous C4 (Maize) Archaeological skeletal histologic change through time:

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
452 References

comparison of an archaic Native American Burton JH, Wright LE. 1995. Nonlinearity in the
population with modern populations. relationship between bone Sr/Ca and diet:
Anatomical Record 226:307–313. paleodietary implications. American Journal of
Burr DB, Schaffler MB, Yang KH, et al. 1989. The Physical Anthropology 96:273–282.
effects of altered strain environments on bone Bush HM, Stirland A. 1988. Osteological evidence
tissue kinetics. Bone 10:215–221. for decapitations in two Romano-British
Burrell LL, Maas MC, Van Gerven DP. 1986. cemeteries. In: Šimic D, Sujoldžić A, Turek S,
Patterns of long-bone fracture in two Nubian editors. Abstracts of the 12th International
cemeteries. Human Evolution 1:495–506. Congress of Anthropological and Ethnological
Burt BA. 1993. Relative consumption of sucrose Sciences. Collegium Anthropologicum 12:344.
and other sugars: has it been a factor in Buxton LHD. 1938. Platymeria and platycnemia.
reduced caries experience? Caries Research Journal of Anatomy 73:31–36.
27(S1):56–63. Buzon MR. 2006. Health of the non-elites at
Burt BA, Eklund SA, Landis JR, et al. 1982. Diet Tombos: nutritional and disease stress in New
and dental health, a study of relationships. Kingdom Nubia. American Journal of Physical
National Center for Health Statistics Series Anthropology 130:26–37.
11:225. Buzon MR. 2011. Nubian identity in the Bronze
Burt BA, Eklund SA, Morgan KJ, et al. 1988. The Age: patterns of cultural and biological
effects of sugars intake and frequency of variation. Bioarchaeology of the Near East
ingestion on dental caries increment in a 5:19–40.
three-year longitudinal study. Journal of Buzon MR. 2012. The bioarchaeological approach
Dental Research 67:1422–1429. to paleopathology. In: Grauer AL, editor.
Burt BA, Ismail AI. 1986. Diet, nutrition, and food A Companion to Paleopathology. Chichester,
cariogenicity. Journal of Dental Research 65 UK: Wiley-Blackwell; p. 58–75.
(Special Issue):1475–1484. Buzon MR, Judd MA. 2008. Investigating health at
Burt BA, Pai S. 2001. Sugar consumption and Kerma: sacrificial versus nonsacrificial
caries risk: a systematic review. Journal of individuals. American Journal of Physical
Dental Education 65:1017–1023. Anthropology 136:93–99.
Burton JH. 1996. Trace elements in bone as Buzon MR, Richman R. 2007. Traumatic injuries
paleodietary indicators. In: Orna MV, editor. and imperialism: the effects of Egyptian
Archaeological Chemistry: Organic, Inorganic, colonial strategies at Tombos in Upper Nubia.
and Biochemical Analysis. Washington, DC: American Journal of Physical Anthropology
American Chemical Society; p. 327–333. 133:783–791.
Burton JH. 2008. Bone chemistry and trace Buzon MR, Simonetti A, Creaser RA. 2007.
element analysis. In: Katzenberg MA, Saunders Migration in the Nile Valley during the New
SR, editors. Biological Anthropology of the Kingdom period: a preliminary strontium
Human Skeleton. Hoboken, NJ: Wiley-Liss; isotope study. Journal of Archaeological
p. 443–460. Science 34:1391–1401.
Burton JH, Price TD. 1990a. Paleodietary Byrd JE, Jantz RL. 1994. Osteological evidence for
applications of barium values in bone. In: distinct social groups at the Leavenworth site.
Pernicka E, Wagner GA, editors. Proceedings of In: Owsley DW, Jantz RL, editors. Skeletal
the 27th International Symposium on Biology in the Great Plains: Migration,
Archaeometry. Basel, Switzerland: Berkhauser Warfare, Health, and Subsistence. Washington,
Verlag AG; p. 1–9. DC: Smithsonian Institution Press;
Burton JH, Price TD. 1990b. The ratio of barium to p. 303–308.
strontium as a paleodietary indicator of Byrne KB, Parris DC. 1987. Reconstruction of the
consumption of marine resources. Journal of diet of the Middle Woodland Amerindian
Archaeological Science 17:547–557. population at Abbott Farm by bone

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 453

trace-element analysis. American Journal of Canci A, Minozzi S, Borgognini Tarli S. 1996. New
Physical Anthropology 74:373–384. evidence of tuberculosis spondylitis from
Cabana GS, Hunley K, Kaestle FA. 2008. Neolithic Liguria (Italy). International Journal
Population continuity or replacement? A novel of Osteoarchaeology 6:497–501.
computer simulation approach and its Cardoso HFV. 2007. Environmental effects on
application to the Numic expansion (Western skeletal versus dental development: using a
Great Basin, USA). American Journal of documented subadult skeletal sample to test a
Physical Anthropology 135:438–447. basic assumption in human osteological
Cabrera Castro R. 1993. Human sacrifice at the research. American Journal of Physical
Temple of the Feathered Serpent: recent Anthropology 132:223–233.
discoveries at Teotihuacan. In: Berrin K, Cardoso HFV, Gomes JEA. 2009. Trends in adult
Pasztory E, editors. Teotihuacan: Art from the stature of peoples who inhabited the modern
City of the Gods. New York, NY: Thames & Portuguese territory from the Mesolithic to the
Hudson; p. 101–107. late 20th century. International Journal of
Cabrera Castro R, Sugiyama S, Cowgill GL. 1991. Osteoarchaeology 19:711–725.
The Templo de Quetzalcoatl Project at Carlos JP, Gittelsohn AM. 1965. Longitudinal
Teotihuacan: a preliminary report. Ancient studies of the natural history of caries. II. A life
Mesoamerica 2:77–92. table study of caries incidence in the
Cadien JD, Harris EF, Jones WP, Mandarino LJ. permanent teeth. Archives of Oral Biology
1974. Biological lineages, skeletal populations, 10:739–751.
and microevolution. Yearbook of Physical Carlson DS. 1976. Patterns of morphological
Anthropology 18:194–201. variation in the human midface and upper face.
Caffey J. 1931. Clinical and experimental lead In: McNamara J Jr., editor. Factors Affecting
poisoning: some roentgenologic and anatomic the Growth of the Midface. University of
changes in growing bone. Radiology Michigan Center for Human Growth and
17:957–983. Development, Craniofacial Growth Series
Calcagno JM. 1986a. Dental reduction in Monograph, No. 6; p. 277–299.
post-Pleistocene Nubia. American Journal Carlson DS. 2005. Theories of craniofacial growth
of Physical Anthropology 70:349–363. in the postgenomic era. Seminars in
Calcagno JM. 1986b. Odontometrics and Orthodontics 11:172–183.
biological continuity in the Meroitic, X-Group, Carlson DS, Armelagos GJ, Van Gerven DP. 1974.
and Christian phases of Nubia. Current Factors influencing the etiology of cribra
Anthropology 27:66–69. orbitalia in prehistoric Nubia. Journal of
Calcagno JM. 1989. Mechanisms of Human Dental Human Evolution 3:405–410.
Reduction: A Case Study from Post-Pleistocene Carlson DS, Armelagos GJ, Van Gerven DP. 1976.
Nubia. University of Kansas Publications in Patterns of age-related cortical bone loss
Anthropology, No. 18. (osteoporosis) within the femoral diaphysis.
Calcagno JM, Gibson KR. 1991. Selective Human Biology 48:295–314.
compromise: evolutionary trends and Carlson DS, Van Gerven DP. 1977. Masticatory
mechanisms in hominid tooth size. In: Kelley function and post-Pleistocene evolution in
MA, Larsen CS, editors. Advances in Dental Nubia. American Journal of Physical
Anthropology. New York, NY: Wiley-Liss; Anthropology 46:495–506.
p. 59–76. Carlson DS, Van Gerven DP. 1979. Diffusion,
Cameron N, Tobias P, Fraser W, Nagdee M. 1990. biological determinism, and biocultural
Search for secular trends in calvarial diameters, adaptation in the Nubian Corridor. American
cranial base height, indexes, and capacity in Anthropologist 81:561–580.
South African Negro crania. American Journal Carmichael AG. 1993. Leprosy. In: Kiple KF, editor.
of Human Biology 2:53–61. The Cambridge World History of Human

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
454 References

Disease. Cambridge, UK: Cambridge University Chamay A, Tschantz P. 1972. Mechanical


Press; p. 834–839. influences in bone remodeling: experimental
Carneiro RL. 1990. Chiefdom-level warfare as research on Wolff’s Law. Journal of
exemplified in Fiji and the Cauca Valley. In: Biomechanics 5:173–180.
Haas J, editor. The Anthropology of War. New Chamberlain AT. 2006. Demography in
York, NY: Cambridge University Press; Archaeology. Cambridge, UK: Cambridge
p. 190–211 University Press.
Carnese FR, Mendisco F, Keyser C, et al. 2010. Chamberlain AT, Parker Pearson M. 2001. Earthly
Paleogenetical study of pre-Columbian Remains: The History and Science of Preserved
samples from Pampa Grande (Salta, Bodies. New York, NY: Oxford University
Argentina). American Journal of Physical Press.
Anthropology 141:452–462. Chapman FH. 1972. Vertebral osteophytosis in
Carroll ST. 1988. Wrestling in ancient Nubia. prehistoric populations of central and southern
Journal of Sport History 15:121–137. Mexico. American Journal of Physical
Carson A. 2006. Maximum-likelihood variance Anthropology 36:31–37.
components analysis of heritabilities of Chapman NEM. 1997. Evidence for Spanish
cranial nonmetric traits. Human Biology influence on activity induced musculoskeletal
78:383–402. stress markers at Pecos Pueblo. International
Carson J. 1985. Colonial Virginia Cookery: Journal of Osteoarchaeology 7:497–506.
Procedures, Equipment, and Ingredients in Chapman R, Kinnes I, Randsborg K. 1981. The
Colonial Cooking. Williamsburg, VA: Colonial Archaeology of Death. Cambridge, UK:
Williamsburg Foundation. Cambridge University Press.
Carter DR, Beaupré GS. 2001. Skeletal Function Chard CS. 1974. Northeast Asia in Prehistory.
and Form: Mechanobiology of Skeletal Madison, WI: University of Wisconsin Press.
Development, Aging, and Regeneration. Chari S, Lavallee MN, editors. 2013.
Cambridge, UK: Cambridge University Press. Accomplishing NAGPRA: Perspectives on the
Carvalho ML, Marques AF, Lima MT, Reus U. Intent, Impact, and Future of the Native
2004. Trace element distribution and post- American Graves Protection and Repatriation
mortem intake in human bones from Middle Act. Corvallis, OR: Oregon State University
Age by total reflection X-ray fluorescence. Press.
Spectrochimica Acta Part B: Atomic Charles RH. 1893. The influence of function, as
Spectroscopy 59:1251–1257. exemplified in the morphology of the lower
Cassidy CM. 1984. Skeletal evidence for extremity of the Punjabi. Journal of Anatomy
prehistoric subsistence adaptation in the and Physiology 28:1–18.
central Ohio River valley. In: Cohen MN, Chase AF, Chase DZ. 2001. The Royal Court of
Armelagos GJ, editors. Paleopathology at the Caracol, Belize: its palaces and people. In:
Origins of Agriculture. Orlando, FL: Academic Inomata T, Houston SD, editors. Royal Courts
Press; p. 307–345. of the Ancient Maya, Volume 2. Boulder, CO:
Cavalli-Sforza LL, Menozzi P, Piazza A. 1996. The Westview; p. 102–137.
History and Geography of Human Genes. Chatters JC, Kennett DJ, Asmerom Y, et al. 2014.
Princeton, NJ: Princeton University Press. Late Pleistocene human skeleton and mtDNA
Chagnon N. 1992. Yanomamö, Fourth Edition. link Paleoamericans and modern Native
Fort Worth, TX: Harcourt Brace Janovich. Americans. Science 344:750–754.
Chakrabarti D. 1980. Early agriculture and the Chenery C, Eckardt H, Muldner G. 2011.
development of towns in India. In: Sherrat A, Cosmopolitan Catterick? Isotopic evidence for
editor. The Cambridge Encyclopedia of population mobility on Rome’s northern
Archaeology. Cambridge, UK: Cambridge frontier. Journal of Archaeological Science
University Press; p. 162–167. 38:1525–1536.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 455

Chesson MS, editor. 2001. Social Memory, Identity WA: Washington State University Press;
and Death: Anthropological Perspectives on p. 69–74.
Mortuary Rituals. Archeological Papers of the Cho H, Stout SD. 2003. Bone remodeling and
American Anthropological Association, No. 10. age-associated bone loss in the past: a
Chesterman JT. 1983. The human skeletal remains. histomorphometric analysis of the Imperial
In: Hedges JW, editor. Isbister: A Chambered Roman skeletal population of Isola Sacra. In:
Tomb in Orkney. British Archaeological Report, Agarwal SC, Stout SD, editors. Bone Loss and
British Series, No. 115; p. 73–132. Osteoporosis: An Anthropological Perspective.
Cheverud JM, Buikstra JE. 1981a. Quantitative New York, NY: Kluwer Academic Plenum;
genetics of skeletal nonmetric traits in the p. 207–228.
rhesus macaques of Cayo Santiago. II. Cho H, Stout SD, Bishop TA. 2006. Cortical bone
Phenotypic, genetic and environmental remodeling rates in a sample of African
correlations between traits. American Journal American and European American descent
of Physical Anthropology 54:51–58. groups from the American Midwest:
Cheverud JM, Buikstra JE. 1981b. Quantitative comparisons of age and sex in ribs. American
genetics of skeletal nonmetric traits in the Journal of Physical Anthropology 130:
rhesus macaques on Cayo Santiago. I. Single 214–226.
trait heritabilities. American Journal of Choy K, Ok-Ryun J, Fuller BT, Richards MP. 2010.
Physical Anthropology 49:43–49. Isotopic evidence of dietary variations and
Cheverud JM, Buikstra JE. 1982. Quantitative weaning practices in the Gaya cemetery at
genetics of skeletal nonmetric traits in the Yeanri, Gimhae, South Korea. American
rhesus macaques of Cayo Santiago. III. Relative Journal of Physical Anthropology 142:74–84.
heritability of skeletal nonmetric and metric Christensen AF. 1998. Odontometric
traits. American Journal of Physical microevolution in the Valley of Oaxaca,
Anthropology 59:151–155. Mexico. Journal of Human Evolution 34:
Cheverud JM, Buikstra JE, Twitchell E. 1979. 333–360.
Relationships between non-metric skeletal Christensen AF, Winter M. 1997. Culturally
traits and cranial size and shape. American modified skeletal remains from the site of
Journal of Physical Anthropology 50:191–198. Huamelulpan, Oaxaca, Mexico. International
Chhem RK, Brothwell DR. 2007. Paleoradiology: Journal of Osteoarchaeology 7:467–480.
Imaging Mummies and Fossils. Berlin, Christopherson KM, Pedersen PO. 1939.
Germany: Springer. Investigations into dental conditions in the
Chhem RK, Rühli FJ. 2004. Paleoradiology of Neolithic period and in the Bronze Age in
mummies, skeletal remains and hominid Denmark. Dental Record 59:575–585.
fossils. Canadian Association of Radiologists Churchill SE. 1994. Human Upper Body Evolution
Journal 55:193–279. in the Eurasian Later Pleistocene. PhD
Chisholm B, Koike H. 1999. Reconstructing Dissertation, University of New Mexico,
prehistoric Japanese diet using stable isotopic Albuquerque, NM.
analysis. In: Omoto K, editor. Interdisciplinary Churchill SE, Smith FH. 2000. Makers of the Early
Perspectives on the Origins of the Japanese. Aurignacian of Europe. Yearbook of Physical
International Research Center for Japanese Anthropology 43:61–115.
Studies, Kyoto; p. 199–222. Cicchitti A. 1993. Pompei: Il Primo Calco
Chisholm B, Koike H, Nakai N. 1992. Carbon Trasparente (Diario di un Scavo). Rome, Italy:
isotopic determination of paleodiet in Japan: L’Erma di Bretschneider.
marine versus terrestrial resources. In: Aikens Ciranni R, Fornaciari G. 2003. Luigi Boccherini
CM, Nai Rhee S, editors. Pacific Northeast Asia and the Barocco cello: an 18th century striking
in Prehistory: Hunter-Fisher-Gatherers, case of occupational disease. International
Farmer, and Sociopolitical Elites. Pullman, Journal of Osteoarchaeology 13:294–302.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
456 References

Clabeaux MS. 1977. Health and disease in the Coale AJ. 1972. The Growth and Structure of
population of an Iroquois ossuary. Yearbook of Human Populations. Princeton, NJ: Princeton
Physical Anthropology 20:359–370. University Press.
Clark AL, Tayles N, Halcrow SE. 2014. Aspects of Coatsworth JH. 1996. Welfare. The American
health in prehistoric mainland Southeast Asia: Historical Review 101:1–12.
indicators of stress in response to the Cockburn TA, Cockburn E, Reyman TA, editors.
intensification of rice agriculture. American 1998. Mummies, Disease and Ancient Cultures.
Journal of Physical Anthropology 153:484–495. Cambridge, UK: Cambridge University Press.
Clark GA. 1988. New method for assessing Coggon D, Kellingray S, Inskip H, et al. 1998.
changes in growth and sexual dimorphism in Osteoarthritis of the hip and occupational
paleoepidemiology. American Journal of lifting. American Journal of Epidemiology
Physical Anthropology 77:105–116. 147:523–528.
Clark GA, Hall NR, Armelagos GJ, et al. 1986 Poor Cohen MN. 1989. Health and the Rise of
growth prior to early childhood: decreased Civilization. New Haven, CT: Yale University
health and life-span in the adult. American Press.
Journal of Physical Anthropology 70:145–160. Cohen MN. 1998. The emergence of health and
Clark JGD. 1972. Star Carr: A Case Study in social inequalities in the archaeological record.
Bioarchaeology. Reading, MA: Addison- In: Strickland SS, Shetty PS, editors. Human
Welsey. Biology and Social Inequality. Cambridge, UK:
Clarke N. 1993. Periodontitis in dry skulls. Dental Cambridge University Press; p. 249–271.
Anthropology Newsletter 7:1–4. Cohen MN, Armelagos GJ, editors. 1984.
Clarke NB, Carey SE, Srikandi W, Hirsch RS, Paleopathology at the Origins of Agriculture.
Leppard PI. 1986. Periodontal disease in Orlando, FL: Academic Press. (Reprinted,
ancient populations. American Journal of University Press of Florida, Gainesville, FL,
Physical Anthropology 71:173–183. 2013.)
Clarke NG, Hirsch RS. 1991. Physiological, pulpal, Cohen MN, Baum BJ, Garn SM, Osorio CH, Nagy
and periodontal factors influencing alveolar HM. 1979. Crown-size reduction in congenital
bone. In: Kelley MA, Larsen CS, editors. defects. In: Dahlberg A, Graber TM, editors.
Advances in Dental Anthropology. New York, Orofacial Growth and Development. The Hague,
NY: Wiley-Liss; p. 241–266. the Netherlands: Mouton; p. 119–126.
Clarke SK. 1980. Early childhood morbidity trends Cohen MN, Crane-Kramer GMM, editors. 2007.
in prehistoric populations. Human Biology Ancient Health: Skeletal Indicators of
52:79–85. Agricultural and Economic Intensification.
Clarke SK, Gindhart PS. 1981. Commonality in Gainesville, FL: University Press of Florida.
peak age of early-childhood morbidity across Cohen MN, O’Connor K, Danforth M, Jacobi K,
cultures and over time. Current Anthropology Armstrong C. 1994. Health and death and Tipu.
22:574–575. In: Larsen CS, Milner GR, editors. In the Wake
Clarkson JE, Worthington HV. 1993. Association of Contact: Biological Responses to Conquest.
between untreated caries and age, gender and New York, NY: Wiley-Liss; p. 121–133.
dental attendance in adults. Community Cole G, Waldron T. 2011. Apple Down 152: a
Dentistry and Oral Epidemiology 21:126–128. putative case of syphilis from sixth century AD
Claussen BF. 1982. Chronic hypertrophy of the Anglo-Saxon England. American Journal of
ulna in the professional rodeo cowboy. Clinical Physical Anthropology 144:72–79.
Orthopaedics and Related Research 164:45–47. Cole GG, Hill MC, Ensor HB. 1982.
Clement AF, Hillson SW. 2012. Intrapopulation Bioarchaeological comparisons of the Late
variation in macro tooth wear patterns – a case Miller III and Summerville I phases in the
study from Igloolik, Canada. American Journal Gainesville Lake Area. In: Jenkins NJ, editor.
of Physical Anthropology 149:517–524. Archaeology of the Gainesville Lake Area:

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 457

Synthesis. University of Alabama, Office of and Interpretation. Salt Lake City, UT:
Archaeological Research, Report of University of Utah Press; p. 55–83.
Investigations, No. 23; p. 187–258. Comuzzie AG, Steele DG. 1989. Enlarged occlusal
Cole MS, Cole TM III. 1994. Metric variation in the surfaces on first molars due to severe attrition
supraorbital region in northern Plains Indians. and hypercementosis: examples from
In: Owsley DW, Jantz RL, editors. Skeletal prehistoric coastal populations of Texas.
Biology of the Great Plains: Migration, American Journal of Physical Anthropology
Warfare, Health, and Subsistence. Washington, 78:9–15.
DC: Smithsonian Institution Press; Condon K, Rose JC. 1992. Intertooth and
p. 209–217. intratooth variability in the occurrence of
Cole TM III. 1994. Size and shape of the femur and developmental defects. In: Goodman AH,
tibia in Northern Plains Indians. In: Owsley Capasso LL, editors. Recent Contributions to the
DW, Jantz RL, editors. Skeletal Biology of the Study of Enamel Developmental Defects.
Great Plains: Migration, Warfare, Health, and Journal of Paleopathology, Monographic
Subsistence. Washington, DC: Smithsonian Publications, No. 2; p. 61–77.
Institution Press; p. 219–233. Condon KW. 1981. The correspondence of
Collins HB. 1951. The Origin and Antiquity of the developmental enamel defects between the
Eskimo. Annual Report of the Smithsonian mandibular canine and first premolar.
Institution. Washington, DC: Smithsonian American Journal of Physical Anthropology
Institution Press. 54:211.
Coltrain JB, Hayes MG, O’Rourke DH. 2006. Condon RG. 1983. Seasonal photoperiodism,
A radiometric evaluation of Hrdlička’s Aleutian activity rhythms, and disease susceptibility in
replacement hypothesis: population continuity the central Canadian Arctic. Arctic
and morphological change. Current Anthropology 20:33–48.
Anthropology 47:537–548. Conklin BA. 2001. Consuming Grief:
Coltrain JB, Janetski JC. 2013. The stable andraio- Compassionate Cannibalism in an Amazonian
isotope chemistry of southeastern Utah Society. Austin, TX: University of Texas Press.
Basketmaker II burials: dietary analysis using Conlee C. 2007. Decapitation and rebirth: a
the linear mixing model SISUS, age and sex headless burial from Nazca, Peru. Current
patterning, geolocation and temporal Anthropology 48:438–445.
patterning. Journal of Archaeological Science Conlee CA, Buzon MR, Gutierrez AN, Simonetti A,
40:4711–4730. Creaser RA. 2009. Identifying foreigners versus
Coltrain JB, Janetski JC, Carlyle SW. 2007. The locals in a burial population from Nasca, Peru:
stable- and radio-isotope chemistry of Western an investigation using strontium isotope
Basketmaker burials: implications for the analysis. Journal of Archaeological Science
Early Puebloan diets and origins. American 36:2755–2764.
Antiquity 72:301–321. Conner MD. 1984. Population Structure and
Coltrain JB, Leavitt SW. 2002. Climate and diet in Biological Variation in the Late Woodland of
Fremont prehistory: economic variability and West-Central Illinois. PhD Dissertation,
abandonment of maize agriculture in the Great University of Chicago, Chicago, IL.
Salt Lake Basin. American Antiquity Conner MD. 1990. Population structure and
67:453–485. skeletal variation in the Late Woodland of
Coltrain JB, Stafford TW Jr. 1999. Stable carbon West-Central Illinois. American Journal of
isotopes and Great Salt Lake wetlands diet: Physical Anthropology 82:21–43.
towards an understanding of the Great Basin Constandse-Westermann TS. 1982. A skeleton
Formative. In: Hemphill BE, Larsen CS, editors. found in a Roman well at Velsen (Province
Prehistoric Lifeways in the Great Basin North Holland, the Netherlands). Helinium
Wetlands: Bioarchaeological Reconstruction 22:135–169.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
458 References

Constandse-Westermann TS, Newell RR. 1984. Buikstra JE, editor. A Life in Science: Papers in
Mesolithic trauma: demographical and Honor of J. Lawrence Angel. Center for
chronological trends in western Europe. In: American Archeology, Scientific Papers, No. 6;
Haneveld GT, Perizonius WRK, editors. p. 64–86.
Proceedings of the Fourth European Meeting of Cook DC. 1994. Dental evidence for congenital
the Paleopathology Association, Middelburg/ syphilis (and its absence) before and after the
Antwerpen. Utrecht, the Netherlands: Elinkwjk; conquest of the New World. In: Dutour O,
p. 70–76. Palfi G, Berato J, Brun J-P, editors. L’Origine de
Constandse-Westermann TS, Newell RR. 1989. la Syphilis en Europe: Avant ou Après 1493?
Limb lateralization and social stratification in Paris, France: Éditions Errance; p. 169–175.
western European Mesolithic societies. In: Cook DC. 2006. The old physical anthropology and
Hershkovitz I, editor. People and Culture in the New World: a look at the accomplishments
Change. British Archaeological Reports, of an antiquated paradigm. In: Buikstra JE,
International Series, No. 508; p. 405–433. Beck LA, editors. Bioarchaeology: The
Constandse-Westermann TS, Newell RR. 1990. Contextual Analysis of Human Remains.
A diachronic and chronological analysis of Burlington, VT: Academic Press; p. 27–71.
lateralization manifestations in the western Cook DC. 2007. Maize and Mississippians in the
European Mesolithic skeletal sample: a novel American Midwest. In: Cohen MN, Crane-
approach to the assessment of social Kramer GMM, editors. Ancient Health: Skeletal
complexity. In: Vermeersch PM, Van Peer P, Indicators of Agricultural and Economic
editors. Contributions to the Mesolithic in Intensification. Gainesville, FL: University
Europe. Leuven, Belgium: Leuven University Press of Florida; p. 10–19.
Press; p. 95–120. Cook DC, Buikstra JE. 1979. Health and
Cook DC. 1976. Pathologic States and Disease differential survival in prehistoric populations:
Process in Illinois Woodland Populations: An prenatal dental defects. American Journal of
Epidemiologic Approach. PhD Dissertation, Physical Anthropology 51:649–664.
University of Chicago, Chicago, IL. Cook DC, Buikstra JE, DeRousseau CJ, Johanson
Cook DC. 1979. Subsistence base and health in DC. 1983. Vertebral pathology in the Afar
prehistoric Illinois Valley: evidence from the australopithecines. American Journal of
human skeleton. Medical Anthropology Physical Anthropology 60:83–101.
3:109–124. Cook DC, Powell ML. 2012. Treponematosis: past,
Cook DC. 1981a. Mortality, age structure, and present, and future. In: Grauer AL, editor.
status in the interpretation of stress indicators A Companion to Paleopathology. Chichester,
in prehistoric skeletons: a dental example from UK: Wiley-Blackwell; p. 472–491.
the lower Illinois Valley. In: Chapman R, Cook M, Molto E, Anderson C. 1989.
Kinnes I, Randsborg K, editors. The Fluorochrome labelling in Roman period
Archaeology of Death. Cambridge, UK: skeletons from Dakhleh Oasis, Egypt. American
Cambridge University Press; p. 133–144. Journal of Physical Anthropology 80:137–143.
Cook DC. 1981b. Koniag Eskimo tooth ablation: Cook RA. 2005. Reconstructing perishable
Was Hrdlička right after all? Current architecture: prospects and limitations of a Fort
Anthropology 22:159–163. Ancient example. North American
Cook DC. 1984. Subsistence and health in the Archaeologist 26:357–388.
lower Illinois Valley: osteological evidence. In: Cook RA. 2012. Dogs of war: potential social
Cohen, MN, Armelagos GJ, editors. institutions of conflict, healing, and death in a
Paleopathology at the Origins of Agriculture. Fort Ancient village. American Antiquity
Orlando, FL: Academic Press; p. 235–269. 77:498–523.
Cook DC. 1990. Epidemiology of circular caries: a Cook RA, Aubry BS. 2014. Aggregation,
perspective from prehistoric skeletons. In: interregional interaction, and postmarital

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 459

residence patterning: a study of biological Corruccini RS. 1984. An epidemiologic transition


variation in the late prehistoric middle Ohio in dental occlusion in world populations.
Valley. American Journal of Physical American Journal of Orthodontics 82:371–376.
Anthropology 154:270–278. Corruccini RS. 1991. Anthropological aspects of
Cook RA, Schurr MR. 2009. Eating between the orofacial and occlusal variations and
lines: Mississippian migration and stable anomalies. In: Kelley MA, Larsen CS, editors.
carbon isotope variation in Fort Ancient Advances in Dental Anthropology. New York,
populations. American Anthropologist NY: Wiley-Liss; p. 295–323.
111:344–359. Corruccini RS, Beecher RM. 1982. Occlusal
Cooke RG, Norr L, Piperno DR. 1996. Native variation related to soft diet in a nonhuman
Americans and the Panamanian landscape. In: primate. Science 218:74–76.
Reitz EJ, Newsom LA, Scudder SJ, editors. Case Corruccini RS, Beecher RM. 1984. Occlusofacial
Studies in Environmental Archaeology. New morphological integration lowered in baboons
York, NY: Plenum Press; p. 103–126. raised on soft diet. Journal of Craniofacial
Coon CS, Garn SM, Birdsell JB. 1950. Races: Genetics and Developmental Biology
A Study of the Problems of Race Formation in 4:135–142.
Man. Springfield, IL: Charles C. Thomas. Corruccini RS, Choudhury AFH. 1985. Dental
Cooper C, Cawley M, Bhalla A, et al. 1995. occlusal variation among rural and urban
Childhood growth, physical activity, and peak Bengali youths. Human Biology 58:61–66.
bone mass in women. Journal of Bone and Corruccini RS, Handler JS. 1980.
Mineral Research 10:940–947. Temporomandibular joint size decrease in
Cooper C, McAlindon T, Coggon D, Egger P, American Blacks: evidence from Barbados.
Dieppe P. 1994. Occupational activity and Journal of Dental Research 59:1528.
osteoarthritis of the knee. Annals of the Corruccini RS, Handler JS, Jacobi KP. 1985.
Rheumatic Diseases 53:90–93. Chronological distribution of enamel
Cooper JF. 1919. The Last of the Mohicans. New hypoplasias and weaning in a Caribbean slave
York, NY: Charles Scribner’s Sons. population. Human Biology 57:699–711.
Coppa A, Bondioli L, Cucina A, et al. 2006. Early Corruccini RS, Kaul SS, Chopra SRK, et al. 1983.
Neolithic tradition of dentistry. Nature Epidemiological survey of occlusion in north
440:755. India. British Journal of Orthodontics
Coppa A, Cucina A, Chiarelli B, Calderon FL, 10:44–47.
Mancinelli D. 1995. Dental anthropology and Corruccini RS, Potter RHY, Dahlberg AA. 1981.
paleodemography of the Precolumbian Changing occlusal variation in Pima Amerinds.
populations of Hispaniola from the third American Journal of Physical Anthropology
millennium B.C. to the Spanish conquest. 62:317–324.
Human Evolution 10:153–167. Corruccini RS, Shimada I. 2002. Dental relatedness
Corbett ME, Moore WJ. 1976. The distribution of corresponding to mortuary patterning at Huaca
dental caries in ancient British populations: the Loro, Peru. American Journal of Physical
19th century. Caries Research 10:401–412. Anthropology 117:113–121.
Cornejo OE, Lefébure T, Pavinski Bitar PD, et al. Corruccini RS, Townsend GC, Brown T. 1990.
2013. Evolutionary and population genomics Occlusal variation in Australian Aboriginals.
of the cavity causing bacteria Streptococcus American Journal of Physical Anthropology
mutans. Molecular Biology and Evolution 82:257–265.
30:881–893. Corruccini RS, Townsend GC, Schwerdt W. 2005.
Corruccini RS. 1972. The biological relationships Correspondence between enamel hypoplasia
of some prehistoric and historic Pueblo and odontometric biolateral asymmetry in
populations. American Journal of Physical Australian twins. American Journal of Physical
Anthropology 37:373–388. Anthropology 126:177–182.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
460 References

Corruccini RS, Whitley LD. 1981. Occlusal Cox KJ, Bentley RA, Tayles N, et al. 2011. Intrinsic
variation in a rural Kentucky community. or extrinsic population growth in Iron Age
American Journal of Orthodontics northeast Thailand? The evidence from isotopic
79:250–262. analysis. Journal of Archaeological Science
Corti MC, Rigon C. 2003. Epidemiology of 38:665–671.
osteoarthritis: prevalence, risk factors and Coyston S, White CD, Schwarcz HP. 1999. Dietary
functional impact. Aging Clinical and carbonate analysis of bone and enamel for two
Experimental Research 15:359–363. sites in Belize. In: White CD, editor.
Costa D, Steckel RH. 1997. Long-term trends in Reconstructing Ancient Maya Diet. Salt Lake
health, welfare, and economic growth in the City, UT: University of Utah Press; p. 221–243.
United States. In: Steckel RH, Floud R, editors. Craig OE, Biazzo M, O’Connell TC, et al. 2009. Stable
Health and Welfare during Industrialization. isotopic evidence for diet at the Imperial Roman
Chicago, IL: University of Chicago Press; coastal site of Velia (1st and 2nd centuries AD) in
p. 47–90. southern Italy. American Journal of Physical
Costa RL Jr. 1980. Age, sex, and antemortem loss Anthropology 139:572–583.
of teeth in prehistoric Eskimo samples from Crews DE, Bogin B. 2010. Growth, development,
Point Hope and Kodiak Island, Alaska. senescence, and aging: a life history
American Journal of Physical Anthropology perspective. In: Larsen CS, editor.
53:579–587. A Companion to Biological Anthropology.
Costa RL Jr. 1982. Periodontal disease in the Malden, MA: Wiley-Blackwell; p. 124–152.
prehistoric Ipiutak and Tigara skeletal remains Crist TA. 1995. Bone chemistry analysis and
from Point Hope, Alaska. American Journal of documentary archaeology: dietary patterns of
Physical Anthropology 59:97–110. enslaved African Americans in the South
Courville CB. 1962. Forensic neuropathology: IV. Carolina Low Country. In: Grauer AL, editor.
Significance of traumatic extracranial and Bodies of Evidence: Reconstructing History
cranial lesions. Journal of Forensic Sciences through Skeletal Analysis. New York, NY:
7:303–322. Wiley-Liss; p. 197–219.
Courville CB. 1965. War wounds of the cranium in Croft P, Coggon D, Cruddas M, Cooper C. 1992.
the Middle Ages: I. As disclosed in the skeletal Osteoarthritis of the hip: an occupational
material from the Battle of Wisby (1361 A.D.). disease in farmers. British Medical Journal
Bulletin of the Los Angeles Neurological Society 304:1269–1272.
30:27–33. Croft P, Cooper C, Wickham C, Coggon D. 1992.
Coville FV. 1892. The Panamint Indians of Osteoarthritis of the hip and occupational
California. American Anthropologist activity. Scandinavian Journal of Work,
5:351–361. Environment and Health 18:59–63.
Cowgill LW. 2010. The ontogeny of Holocene and Crognier E. 1981. Climate and anthropometric
Late Pleistocene human postcranial strength. variations in Europe and the Mediterranean
American Journal of Physical Anthropology area. Annals of Human Biology 8:99–107.
141:16–37. Crompton RH, Savage R, Spears IR. 1998. The
Cowgill LW, Hager LD. 2007. Variation in the mechanics of food reduction in Tarsius
development of postcranial robusticity: an bancanus: hard-object feeder, soft-object
example from Çatalhöyük, Turkey. feeder or both? Folia Primatologica 69:41–59.
International Journal of Osteoarchaeology Crooks DL. 1994. Growth status of school-age
17:235–252. Mayan children in Belize, Central America.
Cowin SC. 1989. The mechanical properties of American Journal of Physical Anthropology
cancellous bone. In: Cowin SC, editor. Bone 93:217–227.
Mechanics. Boca Raton, FL: CRC Press; Crooks DL. 1995. American children at risk:
p. 129–157. poverty and its consequences for children’s

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 461

health, growth, and school achievement. Cutress TW, Powell RN, Ball ME. 1982. Differing
Yearbook of Physical Anthropology 38:57–86. profiles of periodontal disease in two similar
Crooks DL. 1999. Child growth and nutritional South Pacific island populations. Community
status in a high-poverty community in eastern Dentistry and Oral Epidemiology 10:
Kentucky. American Journal of Physical 193–203.
Anthropology 109:129–142. Cybulski JS. 1974. Tooth wear and material
Crouter SE, DellaValle DM, Haas JD. 2012. culture: precontact patterns in the Tsimshian
Relationship between physical activity, area, British Columbia. Syesis 7:31–35.
physical performance, and iron status in adult Cybulski JS. 1977. Cribra orbitalia, a possible sign
women. Applied Physiology, Nutrition, and of anemia in early historic native populations
Metabolism 37:697–705. of the British Columbia coast. American
Crowder C, Stout S. 2011. editors. Bone Histology: Journal of Physical Anthropology 47:31–40.
An Anthropological Perspective. Boca Raton, Cybulski JS. 1978. An Earlier Population of
FL: CRC Press. Hesquiat Harbour, British Columbia. British
Crubézy E, Legal L, Fabas G, Dabernat H, Ludes B. Columbia Provincial Museum, Cultural
2006. Pathogeny of archaic mycobacteria at Recovery Papers, No. 1.
the emergence of urban life in Egypt (3400 BC). Cybulski JS. 1980. Possible Pre-Columbian
Infection, Genetics, and Evolution 6:13–21. treponematosis on Santa Rosa Island,
Cucina A, Perera Cantillo C, Sierra Sosa T, Tiesler California. Canadian Review of Physical
V. 2011. Carious lesions and maize Anthropology 2:19–25.
consumption among the Prehispanic Maya: an Cybulski JS. 1988. Skeletons in the walls of Old
analysis of a coastal community in northern Quebec. Northeast Historical Archaeology
Yucatan. American Journal of Physical 17:61–84.
Anthropology 145:560–567. Cybulski JS. 1990. Human biology. In: Suttles W,
Cucina A, Tiesler V. 2003. Dental caries and editor. Handbook of North American Indians,
antemortem tooth loss in the northern Peten Volume 7: Northwest Coast. Washington, DC:
area, Mexico: a biocultural perspective on Smithsonian Institution Press; p. 52–59.
social status difference among the classic Cybulski JS. 1992. A Greenville Burial Ground:
Maya. American Journal of Physical Human Remains and Mortuary Elements in
Anthropology 122:1–10. British Columbia Coast Prehistory.
Cunha E, Silva AM. 1997. War lesions from the Archaeological Survey of Canada, Mercury
famous Portuguese medieval battle of Series Paper, No. 146.
Aljubarrota. International Journal of Cybulski JS. 1994. Culture change, demographic
Osteoarchaeology 7:595–599. history, and health and disease on the
Cunha E, Umbelino C, Silva AM, Cardoso F. 2007. Northwest coast. In: Larsen CS, Milner GR,
What can pathology say about the Mesolithic editors. In the Wake of Contact: Biological
and Late Neolithic/Chalcolithic communities? Responses to Conquest. New York, NY: Wiley-
In: Cohen MN, Crane-Kramer GMM, editors. Liss; p. 75–85.
Ancient Health: Skeletal Indicators of Cybulski JS. 2006. Skeletal biology: Northwest
Agricultural and Economic Intensification. coast and Plateau. In: Ubelaker DH, editor.
Gainesville, FL: University Press of Florida; Handbook of North American Indians: 3.
p. 164–175. Environment, Origins, and Population.
Current-Garcia E, Hatfield DB, editors. 1973. Washington, DC: Smithsonian Institution
Shem, Ham & Japheth: The Papers of W.O. Press; p. 532–547.
Tuggle. Athens, GA: University of Georgia Cybulski JS. 2014. Conflict on the northern
Press. Northwest Coast: 2,000 years plus of
Currey JD. 2002. Bones: Structure and Mechanics. bioarchaeological evidence. In: Knüsel C,
Princeton, NJ: Princeton University Press. Smith MJ, editors. The Routledge Handbook of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
462 References

the Bioarchaeology of Human Conflict. London, Danforth ME. 1999. Coming up short: stature and
UK: Routledge; p. 415–452. nutrition among the ancient Maya of the
Dabbs GR. 2011. Health status among prehistoric southern Lowlands. In: White CD, editors.
Eskimos from Point Hope, Alaska. American Reconstructing Ancient Maya Diet. Salt Lake
Journal of Physical Anthropology 146:94–103. City, UT: University of Utah Press; p. 103–117.
Daegling DJ. 2010. Understanding skull function Danforth ME, Jacobi KP, Cohen MN. 1997. Gender
from a mechanicobiological perspective. In: and health among the Colonial Maya of Tipu,
Larsen CS, editor. A Companion to Biological Belize. Ancient Mesoamerica 8:13–22.
Anthropology. Chicester, UK: Wiley-Blackwell; Danforth ME, Jacobi KP, Wrobel G, Glassman S.
p. 501–515. 2007. Health and the transition to horticulture
Daegling DJ, Grine FE. 1999. Terrestrial foraging in the South-Central United States. In: Cohen
and dental microwear in Papio ursinus. MN, Crane-Kramer GMM, editors. Ancient
Primates 40:559–572. Health: Skeletal Indicators of Agricultural and
Da-Gloria P, Larsen CS. 2014. Oral health of the Economic Intensification. Gainesville, FL:
Paleoamericans of Lagoa Santa, Brazil. University Press of Florida; p. 65–79.
American Journal of Physical Anthropology Dangerfield PH. 1994. Asymmetry and growth. In:
154:11–26. Ulijaszek SJ, Mascie-Taylor CGN, editors.
Dahlberg AA. 1951. The dentition of the American Anthropometry: the Individual and the
Indian. In: Laughlin WS, editor. The Physical Population. Cambridge, UK: Cambridge
Anthropology of the American Indian. New University Press; p. 7–29.
York, NY: The Viking Fund; p. 138–176. Danielson DR, Reinhard KJ. 1998. Human dental
Dahlberg AA. 1963. Analysis of the American microwear caused by calcium oxalate
Indian dentition. In: Brothwell DR, editor. phytoliths in prehistoric diet of the lower Pecos
Dental Anthropology. New York, NY: region, Texas. American Journal of Physical
Pergamon Press; p. 147–177. Anthropology 107:297–304.
D’Aiuto F, Parkar M, Brett PM, Ready D, Tonetti Darwin C. 1859. On the Origin of Species. London,
MS. 2004. Gene polymorphisms in pro- UK: John Murray.
inflammatory cytokines are associated with Dasgupta P. 1993. An Inquiry into Well-Being and
systemic inflammation in patients with severe Destitution. New York, NY: Oxford University
periodontal infections. Cytokine 28:29–34. Press.
Daly RM, Saxon L, Turner CH, Robling AG, Bass Davey TF. 1974. The nose in leprosy: steps to a
SL. 2004. The relationship between muscle size better understanding. Leprosy Review
and bone geometry during growth and in 45:97–103.
response to exercise. Bone 34:281–287. Davidson JM, Rose JC, Gutmann MP, et al. 2002.
Dancause KN, Jing Cao X, Veru F, et al. 2012. The quality of African American life in the Old
Prenatal and early postnatal stress exposure Southwest near the turn of the Twentieth
influences long bone length in adult offspring. century. In: Steckel R, Rose JC, editors. The
American Journal of Physical Anthropology Backbone of History: Health and Nutrition in
149:307–311. the Western Hemisphere. New York, NY:
Danforth ME. 1994. Stature change in prehistoric Cambridge University Press; p. 226–277.
Maya of the southern Lowlands. Latin Davies DM. 1972. The Influence of Teeth, Diet, and
American Antiquity 5:206–211. Habits on the Human Face. London, UK:
Danforth ME. 1997. Late Classic Maya health Heinemann.
patterns: evidence from enamel microdefects. Davies TGH, Pedersen PO. 1955. The degree of
In: Whittington SL, Reed DM, editors. Bones of attrition of the deciduous teeth and first
the Maya: Studies of Ancient Skeletons. permanent molars of primitive and urbanised
Washington, DC: Smithsonian Institution Greenland natives. British Dental Journal
Press; p. 127–137. 99:36–43.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 463

Davis WB, Winter PJ. 1980. The effect of abrasion DeMause L. 1974. The History of Childhood.
on enamel and dentine after exposure to New York, NY: Psychohistory Press.
dietary acid. British Dental Journal Demirjian A. 1986. Dentition. In: Falkner F,
148:253–256. Tanner JM, editors. Human Growth:
Deacon HJ, Deacon J. 1999. Human Beginnings in A Comprehensive Treatise, Volume 2.
South Africa: Uncovering the Secrets of the New York, NY: Plenum; p. 269–298.
Stone Age. Walnut Creek, OH: AltaMira Press. Dempsey PJ, Townsend GC. 2001. Genetic
De Azevedo S, Nocera A, Paschetta C, et al. 2011. and environmental contributions to variation
Evaluating microevolutionary models for the in human tooth size. Heredity 86:
early settlement of the New World: the 685–693.
importance of recurrent gene flow with Asia. Dempsey PJ, Townsend GC, Martin NG, Neale NC.
American Journal of Physical Anthropology 1996. Genetic covariance structure of incisor
146:116–129. crown size in twins. Journal of Dental Research
Deguilloux M-F, Leahy R, Pemonge M-H, Rottier 11:1389–1398.
S. 2012. European Neolithization and ancient DeNiro MJ, Epstein S. 1978. Influence of diet on
DNA: an assessment. Evolutionary the distribution of carbon isotopes in animals.
Anthropology 21:24–37. Geochimica et Cosmochimica Acta
Deguilloux M-F, Soler L, Pemonge M-H, et al. 42:295–406.
2011. News from the West: ancient DNA from a DeNiro MJ, Epstein S. 1981. Influence of diet on
French megalithic burial chamber. American the distribution of nitrogen isotopes in animals.
Journal of Physical Anthropology Geochimica et Cosmochimica Acta
144:108–118. 45:341–351.
DeGusta D. 2000. Fijian cannibalism and mortuary DePaola PF, Soparkar PM., Tavares M, Allukian M,
ritual: bioarchaeological evidence from Vunda. Peterson H. 1982. A dental survey of
International Journal of Osteoarchaeology Massachusetts schoolchildren. Journal of
10:76–92. Dental Research 61:1356–1360.
Deiss JJ. 1989. Herculaneum: Italy’s Buried Department of Interior. 2010. Native American
Treasure. Malibu, CA: J. Paul Getty Museum. Graves Protection and Repatriation Act
de la Fuente C, Flores S, Moraga M. 2013. DNA regulations: disposition of culturally
from human ancient bacteria: a novel source of unidentifiable remains. Federal Register
genetic evidence from archaeological dental 75:12378–12405.
calculus. Archaeometry 55:766–778. De Poncins G. 1941. Kablooma. New York, NY:
DeLeon VB. 2007. Fluctuating asymmetry and Reynal.
stress in a Medieval Nubian population. DeStefano F, Anda RF, Kahn HS, Williamson DF,
American Journal of Physical Anthropology Russell CM. 1993. Dental disease and risk of
132:520–534. coronary heart disease and mortality. British
Delgado-Darias T, Velasco-Vázquez J, Arnay-de- Medical Journal 306:688–691.
la-Rosa M, Martín-Rodríguez E, González- Deter CA. 2009. Gradients of occlusal wear in
Reimers E. 2005. Dental caries among the hunter-gatherers and agriculturalists.
Prehispanic population from Gran Canaria. American Journal of Physical Anthropology
American Journal of Physical Anthropology 138:247–254.
128:560–568. Dettwyler KA, Fishman C. 1992. Infant feeding
Delgado-Darias T, Velasco-Vazquez J, Arnay-de- practices and growth. Annual Review of
la-Rosa M, Martin-Rodriguez E, Gonzalez- Anthropology 21:171–204.
Reimers E. 2006. Calculus, periodontal disease Dever WG. 1997. Archaeology and the emergence
and tooth decay among the prehispanic of Early Israel. In: Bartlett JR, editor.
population from Gran Canaria. Journal of Archaeology and Biblical Interpretation.
Archaeological Science 33:663–670. London, UK: Routledge; p. 20–50.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
464 References

Devor EJ. 1987. Transmission of human Dirks W, Reid DJ, Jolly CJ, Phillips-Conroy JE,
craniofacial dimensions. Journal of Brett FL. 2002. Out of the mouths of baboons:
Craniofacial Genetics and Developmental stress, life history, and dental development in
Biology 7:95–106. the Awash National Park Hybrid Zone,
Dewey JR, Armelagos GJ, Bartley MH. 1969. Ethiopia. American Journal of Physical
Femoral cortical involution in three Nubian Anthropology 118:239–252.
archaeological populations. Human Biology Divale WT, Harris M. 1976. Population, warfare,
41:13–28. and the male supremacist complex. American
DeWitte SN. 2012. Sex differences in periodontal Anthropologist 78:521–538.
disease in catastrophic and attritional Dixon RB. 1923. The Racial History of Man. New
assemblages from Medieval London. American York, NY: Scribners.
Journal of Physical Anthropology Dobney K, Goodman A. 1991. Epidemiological
149:405–416. studies of dental enamel hypoplasias in Mexico
DeWitte SN, Bekvalac J. 2010. Oral health and and Bradford: their relevance to archaeological
frailty in the medieval English cemetery of skeletal studies. In: Bush H, Zvelebil M, editors.
St Mary Graces. American Journal of Physical Health in Past Societies: Biocultural
Anthropology 142:341–354. Interpretations of Human Skeletal Remains in
DeWitte SN, Bekvalac J. 2011. The association Archaeological Contexts. British
between periodontal disease and periosteal Archaeological Reports, International Series,
lesions in the St. Mary Graces cemetery, No. 567; p. 81–100.
London, England A.D. 1350–1538. American Dobson M. 2007. Disease: The Extraordinary
Journal of Physical Anthropology Stories behind History’s Deadliest Killers.
146:609–618. London, UK: Quercus.
Dickel DN. 1991. Descriptive Analysis of the Dodo Y. 1986. Metrical and non-metrical analyses
Skeletal Collection from the Prehistoric of Jomon crania from eastern Japan. In:
Manasota Key Cemetery, Sarasota County, Akazawa T, Aikens CM, editors. Prehistoric
Florida (8SO1292). Florida Department of Hunter-Gatherers in Japan: New Research
State, Bureau of Archaeological Research, Methods. University of Tokyo, University
Florida Archaeological Reports, No. 22. Museum Bulletin, No. 27; p. 137–161.
Dickel DN, Schulz PD, McHenry HM. 1984. Central Dolphin AE, Goodman AH, Amarasiriwardena DD.
California: prehistoric subsistence changes and 2005. Variation in elemental intensities among
health. In: Cohen MN, Armelagos GJ, editors. teeth and between pre- and postnatal regions
Paleopathology at the Origins of Agriculture. of enamel. American Journal of Physical
Orlando, FL: Academic Press; p. 439–461. Anthropology 128:878–888.
Dickson GC. 1970. The natural history of Domett K, Tayles N. 2007. Population health from
malocclusion. Dental Practice 20:216–232. the Bronze to the Iron Age in the Mun River
DiGangi EA, Moore MK. 2013. Research Methods valley, Northeastern Thailand. In: Cohen MN,
in Human Skeletal Biology. Oxford, UK: Crane-Kramer GMM, editors. Ancient Health:
Academic Press. Skeletal Indicators of Agricultural and
Dillehay TD, Nuñez AL. 1988. Camelids, caravans, Economic Intensification. Gainesville, FL:
and complex societies in the south-central University Press of Florida; p. 286–299.
Andes. In: Saunders N, de Montmollin O, Domett KM. 2001. Health in Late Prehistoric
editors. Recent Studies in Precolumbian Thailand. Oxford, UK: British Archaeological
Archaeology. British Archaeological Reports; Reports, International Series, No. 946.
p. 603–634. Domett KM, O’Reilly DJW, Buckley HR. 2011.
Dinning TAR. 1949. A study of healed fractures in Bioarchaeological evidence for conflict in Iron
the Australian aboriginal. Medical Journal of Age northwest Cambodia. Antiquity
Australia 12:712–713. 85:441–458.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 465

Domett KM, Oxenham MF. 2010. The demographic Douglas MT. 2006. Subsistence change and dental
profile of the Man Bac cemetery sample. In: health in the people of Non Nok Tha, northeast
Oxenham MF, Matsumura H, Dung NK, editors. Thailand. In: Oxenham M, Tayles N, editors.
Man Bac: The Excavation of a Neolithic Site in Bioarchaeology of Southeast Asia. Cambridge,
Northern Vietnam, The Biology. Canberra, UK: Cambridge University Press; p. 191–219.
Australia: Australian National University Douglas MT, Pietrusewsky M. 2007. Biological
E Press; p. 9–20. consequences of sedentism: agricultural
Domett KM, Tayles N. 2006. Adult fracture intensification in Northeastern Thailand. In
patterns in prehistoric Thailand: a biocultural Cohen MN, Crane-Kramer GMM, editors.
interpretation. International Journal of Ancient Health: Skeletal Indicators of
Osteoarchaeology 16:185–199. Agricultural and Economic Intensification.
Dominick KL, Baker TA. 2004. Racial and ethnic Gainesville, FL: University Press of Florida;
differences in osteoarthritis: prevalence, p. 300–319.
outcomes, and medical care. Ethnicity & Doyle WJ, Johnston O. 1977. On the meaning of
Disease 14:558–566. increased fluctuating dental asymmetry: a
Dongoske KE, Martin DL, Ferguson TJ. 2000. cross-population study. American Journal of
Critique of the claim of cannibalism at Physical Anthropology 46:127–134.
Cowboy Wash. American Antiquity Drake A, Oxenham M. 2012. Disease, climate and
65:179–190. the peopling of the Americas. Historical
Donlon DA. 2000. The value of infracranial Biology DOI:10.1080/08912963. 2012.725728.
nonmetric variation in studies of modern Homo Drake CW, Hunt RJ, Beck JD, Zambon JJ. 1993.
sapiens: an Australian focus. American Journal The distribution and interrelationship of
of Physical Anthropology 113:349–368. Actinobacillus actinomycetemcomitans,
Donoghue HD. 2008. Molecular palaeopathology. Prophyromonas gingivalis, Prevotella
In: Pinhasi R, Mays S, editors. Advances in intermedia, and BANA scores among older
Human Palaeopathology. Chichester, UK: John adults. Journal of Periodontology 64:89–94.
Wiley & Sons, Ltd.; p. 147–176. Drancourt M, Signoli M, Dang LV, et al. 2007.
Donoghue HD, Gladykowska-Rzeczycjka J, Yersinia pestis orientalis in remains of ancient
Marcsik A, Holton J, Spigelman M. 2002. plague patients. Emerging Infectious Diseases
Mycobacterium leprae in archaeological 13:332–333.
samples. In: Roberts CA, Lewis ME, Dreizen S, Currie C, Gilley EJ, Spies TD. 1956.
Manchester K, editors. The Past and Present Observations on the association between
of Leprosy: Archaeological, Historical, nutritive failure, skeletal maturation rate and
Palaeopathological and Clinical Approaches. radiopaque transverse lines in the distal end of
British Archaeological Reports, International the radius in children. American Journal of
Series, No. 1054; p. 271–286. Roentgenology 76:482–487.
Donoghue HD, Marcsik A, Matheson C, et al. 2010. Dreizen S, Spirakis CN, Stone RE. 1964. The
Co-infection of Mycobacterium tuberculosis influence of age and nutritional status on
and Mycobacterium leprae in human “bone scar” formation in the distal end of the
archaeological samples: a possible explanation growing radius. American Journal of Physical
for the historical decline of leprosy. Anthropology 22:295–306.
Proceedings of the Royal Society B Drezner MK. 1995. Osteoporosis types I and II. In:
272:389–394. Anderson JJB, Garner SC, editors. Calcium and
Doran GH, Dickel DN, Ballinger WE Jr, Agee OE, Phosphorus in Health and Disease. Boca Raton,
Laipis PJ. 1986. Anatomical, cellular, and FL: CRC Press; p. 341–354.
molecular analysis of an 8,000-year-old Drinkwater BL. 1994. Does physical activity play a
human brain tissue from Windover role in preventing osteoarthritis? Research
archaeological site. Nature 323:803–806. Quarterly for Exercise and Sport 65:197–206.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
466 References

Driscoll EM, Weaver DS. 2000. Dental health and dentition. American Journal of Physical
Late Woodland subsistence in coastal North Anthropology 134:63–74.
Carolina. In: Lambert PM, editor. Dupras TL, Williams M, De Meyer M, et al. 2010.
Bioarchaeological Studies of Life in the Age of Evidence of amputation as medical treatment
Agriculture: A View from the Southeast. in ancient Egypt. International Journal of
Tuscaloosa, AL: University of Alabama Press; Osteoarchaeology 20:405–423.
p. 168–194. Duray SM. 1990. Deciduous enamel defects and
Droessler J. 1981. Craniometry and Biological caries susceptibility in a prehistoric Ohio
Distance: Biocultural Continuity at the population. American Journal of Physical
Late Woodland–Mississippian Interface. Anthropology 81:27–34.
Center for American Archaeology, Research Duray SM. 1992. Enamel defects and caries
Series, No. 1. etiology: an historical perspective. In:
Drucker D, Bocherens H. 2004. Carbon and Goodman AH, Capasso LL, editors. Recent
nitrogen isotopes as tracers of change in diet Contributions to the Study of Enamel
breadth during Middle Ages and Upper Developmental Defects. Journal of
Paleolithic in Europe. International Journal of Paleopathology, Monographic Publications,
Osteoarchaeology 14:162–177. No. 2; p. 307–319.
Duday H. 2009. The Archaeology of the Dead: Duray SM. 1996a. Dental indicators of stress and
Lectures in Archaeothanatology. Oxford, UK: reduced age at death in prehistoric Native
Oxbow Books. Americans. American Journal of Physical
Dufour DL. 2010. Nutrition, health, and function. Anthropology 99:275–286.
In: Larsen CS, editor. A Companion to Duray SM. 1996b. Stress and mortality: evidence
Biological Anthropology. Malden, MA: Wiley- from the deciduous dentition. American
Blackwell; p. 194–206. Journal of Physical Anthropology Supplement
Dunbar JB. 1969. Dental caries. In: Pelton W, 22:100.
Dunbar J, McMillan R, Moller P, Wolff A, Düring EM. 1997. Specific skeletal injuries
editors. The Epidemiology of Oral Health. observed on the human skeletal remains from
Cambridge, MA: Harvard University Press; the Swedish seventeenth century man-of-war,
p. 1–14. Kronan. International Journal of
Duncan WN. 2012. Biological distance Osteoarchaeology 7:591–594.
analysis in contexts of ritual violence. In: Dutour O. 1986. Enthesopathies (lesions of
Martin DL, Harrod RP, Perez VR, editors. muscular insertions) as indicators of the
The Bioarchaeology of Violence. Gainesville, activities of Neolithic Saharan populations.
FL: University Press of Florida; American Journal of Physical Anthropology
p. 251–275. 71:221–224.
Dupras TL. 2010. The use of stable isotope analysis Dutour O, Buzhilova A. 2014. Palaeolopatho-
to determine infant and young child feeding logical study of Napoleonic mass graves
patterns. In: Moffat T, Prowse T, editors. discovered in Russia. In: Knüsel C, Smith MJ,
Human Diet and Nutrition in Biocultural editors. The Routledge Handbook of the
Perspective: Past Meets Present. Oxford, UK: Bioarchaeology of Human Conflict. London,
Berghahn Books; p. 89–108. UK: Routledge; p. 511–524.
Dupras TL, Schwarcz HP, Fairgrieve SI. 2001. Dutour O, Pálfi G, Bérato J, Brun J-P, editors. 1994.
Infant feeding and weaning practices in Roman L’Origine de la Syphilis en Europe: Avant ou
Egypt. American Journal of Physical Après 1493? Paris, France: Éditions Errance.
Anthropology 115:204–212. Dye DH. 1977. A model for Late Archaic
Dupras TL, Tocheri MW. 2007. Reconstructing subsistence systems in the western Middle
infant weaning histories of Roman period Tennessee Valley during the Bluff Creek phase.
Kellis, Egypt using stable isotope analysis of Tennessee Anthropologist 2:63–80.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 467

Dzierzykray-Rogalski T. 1980. Paleopathology of Interpretations of Human Skeletal Remains in


the Ptolemaic inhabitants of Dakhleh Oasis. Archaeological Contexts. British
Journal of Human Evolution 9:71–74. Archaeological Reports, International Series,
Ebbesen K. 1993. Sacrifices to the power of nature. No. 567; p. 115–127.
In: Hvass S, Storgaard B, editors. Digging into Eisenberg LE. 1991b. Mississippian cultural
the Past: 25 Years of Archaeology in Denmark. terminations in Middle Tennessee: What the
Copenhagen, Denmark: Royal Society of bioarchaeological evidence can tell us. In:
Northern Antiquaries; p. 122–125. Powell ML, Bridges PS, Mires AMW, editors.
Eckardt H, Chenery C, Booth P, et al. 2009. Oxygen What Mean These Bones? Studies in
and strontium isotope evidence for mobility in Southeastern Bioarchaeology. Tuscaloosa, AL:
Roman Winchester. Journal of Archaeological University of Alabama Press; p. 70–88.
Science 36:2816–2825. Eisenstein S. 1978. Spondylolysis. Journal of Bone
Edgerton BC, An K-N, Morrey BF. 1990. Torsional and Joint Surgery 64B:488–494.
strength reduction due to cortical defects in Eli I, Sarnat H, Talmi E. 1989. Effect of the
bone. Journal of Orthopaedic Research birth process on the neonatal line in primary
8:851–855. tooth enamel. Pediatric Dentistry
Edward JB, Benfer RA. 1993. The effects of 11:220–223.
diagenesis on the Paloma skeletal material. In: Elias RW, Hirao Y, Patterson CC. 1982. The
Sandford MK, editor. Investigations of Ancient circumvention of the natural biopurification of
Human Tissue: Chemical Analyses in calcium nutrient pathways by atmosphere
Anthropology. Langhorne, PA: Gordon and inputs of industrial lead. Geochimica et
Breach Science Publishers; p. 183–268. Cosmochimica Acta 46:2561–2580.
Edynak GJ. 1976. Life-styles from skeletal Ellison PT. 1994. Advances in human reproductive
material: a medieval Yugoslav example. In: ecology. Annual Review of Anthropology
Giles E, Friedlander JS, editors. The Measures 23:255–275.
of Man: Methodologies in Biological Ellison PT, O’Rourke MT. 2000. Population growth
Anthropology. Cambridge, MA: Harvard and fertility regulation. In: Stinson S, Bogin B,
University Press; p. 408–432. Huss-Ashmore R, O’Rourke D, editors. Human
Eggan F. 1950. Social organization of the Western Biology: An Evolutionary and Biocultural
Pueblos. Chicago, IL: University of Chicago Perspective. New York, NY: Wiley-Liss;
Press. p. 553–586.
Eggars S, Parks M, Grupe G, Reinhard KJ. 2011. El-Najjar MY. 1976. Maize, malaria and the anemias
Paleoamerican diet, migration and morphology in the Pre-Columbian New World. Yearbook of
in Brazil: archaeological complexity of the Physical Anthropology 20:329–337.
earliest Americans. PLoS ONE 6:e23962. El-Najjar MY. 1979. Human treponematosis and
Eichner ER. 1989. An epidemiologic perspective: tuberculosis: evidence from the New World.
does running cause osteoarthritis? Physician American Journal of Physical Anthropology
and Sportsmedicine 17:147–154. 51:599–618.
Eisenberg LE. 1986. Adaptation in a “Marginal” El-Najjar MY. 1981. A comparative study of facial
Mississippian Population from Middle dimensions. In: Hayes AC, editor. Contributions
Tennessee: Biocultural Insights from to Gran Quivira Archeology. National Park
Paleopathology. PhD Dissertation, New York Service, Publications in Archeology, No. 17;
University, New York City, NY. p. 157–159.
Eisenberg LE. 1991a. Interpreting measures of El-Najjar MY, Andrews J, Moore JG, Bragg DG.
community health during the late prehistoric 1982. Iron deficiency anemia in two prehistoric
period in middle Tennessee: a biocultural American Indian skeletons: a dietary
approach. In: Bush H, Zvelebil M, editors. hypothesis. Plains Anthropologist
Health in Past Societies: Biocultural 27:205–209.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
468 References

El-Najjar MY, Lozoff B, Ryan D. 1975. The Anthropology. Needham Heights, MA: Simon &
paleoepidemiology of porotic hyperostosis in Schuster Custom Publishing; p. 321–338.
the American Southwest: radiological and Erdal YS. 2006. A pre-Columbian case of
ecological considerations. American Journal of congenital syphilis from Anatolia (Nicaea,
Roentgenology, Radium Therapy, and Nuclear 13th century AD). International Journal of
Medicine 125:918–924. Osteoarchaeology 16:16–33.
El-Najjar MY, Robertson AL Jr. 1976. Spongy Erdal YS. 2007. Occlusal grooves in anterior
bones in prehistoric America. Science dentition among Kovuklukaya inhabitants
193:141–143. (Sinop, Northern Anatolia, 10th century AD).
El-Najjar MY, Ryan DJ, Turner CG II, Lozoff B. International Journal of Osteoarchaeology
1976. The etiology of porotic hyperostosis 18:152–166.
among the prehistoric and historic Anasazi Erdal YS, Erdal OD. 2011. A review of trepanations
Indians of southwestern United States. in Anatolia with new cases. International
American Journal of Physical Anthropology Journal of Osteoarchaeology 21:505–534.
44:477–488. Ericson JE. 1985. Strontium isotope
Ember CR. 1978. Myths about hunter-gatherers. characterization in the study of prehistoric
Ethnology 17:439–448. human ecology. Journal of Human Evolution
Ember CR. 1983. The relative decline in women’s 14:503–514.
contribution to agriculture with intensification. Ericson JE. 1989. Some problems and potentials of
American Anthropologist 85:285–304. strontium analysis for human and animal
Ember CR, Ember M. 1997. Violence in the ecology. In: Rundel PW, Ehleringer JR, Nagy
ethnographic record: results of cross-cultural KA, editors. Stable Isotopes in Ecological
research on war and aggression. In: Martin DL, Research. New York, NY: Springer-Verlag;
Frayer DW, editors. Troubled Times: Violence p. 252–259.
and War in the Past. Amsterdam, the Ericson JE, Shirahata H, Patterson CC. 1979.
Netherlands: Gordon and Breach; p. 1–20. Skeletal concentration of lead in ancient
Ensor BE, Irish JD. 1995. Hypoplastic area method Peruvians. New England Journal of Medicine
for analyzing dental enamel hypoplasia. 300:946–951.
American Journal of Physical Anthropology Ericson JE, Smith DR, Flegal AR. 1991. Skeletal
98:507–517. concentrations of lead, cadmium, zinc, and
Enwonwu CO. 1973. Influence of socio-economic silver in ancient North America Pecos Indians.
conditions on dental development in Nigerian Environmental Health Perspectives
children. Archives of Oral Biology 18:95–107. 33:217–224.
Enwonwu CO. 1995. Interface of malnutrition and Ericson JE, West M, Sullivan CH, Krueger HW.
periodontal diseases. American Journal of 1989. The development of maize agriculture in
Clinical Nutrition 61:430S–436S. the Viru Valley, Peru. In: Price TD, editor.
Epker BN, Frost HM. 1966. Periosteal appositional The Chemistry of Prehistoric Human Bone.
bone growth from age two to age seventy in Cambridge, UK: Cambridge University Press;
man: a tetracycline evaluation. Anatomical p. 68–104.
Record 154:573–577. Eriksen MF. 1980. Patterns of microscopic bone
Epker BN, Kelin M, Frost HM. 1965. Magnitude remodeling in three aboriginal American
and location of cortical bone loss in human populations. In: Browman DL, editor. Early
rib with aging. Clinical Orthopaedics Native Americans: Prehistoric Demography,
41:198–203. Economy, and Technology. The Hague, the
Erchak GM. 1996. Family violence. In: Ember CR, Netherlands: Mouton; p. 239–270.
Ember M, Peregrine PN, editors. Research Erikson JD, Lee DV, Bertram JEA. 2000. Fourier
Frontiers in Anthropology: University of North analysis of acetabular shape in Native
Carolina, Chapel Hill, Introduction to American Arikara populations before and after

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 469

acquisition of horses. American Journal of Evensen JP, Øgaard B. 2007. Are malocclusions
Physical Anthropology 113:473–480. more prevalent and severe now? A comparative
Erlandson JM. 1994. Early Hunter-Gatherers of study of Medieval skulls from Norway.
the California Coast. New York, NY: Plenum. American Journal of Orthodontics and
Esclassan R, Grimoud AM, Ruas MP, et al. 2008. Dentofacial Orthopedics 131:710–716.
Dental caries, tooth wear and diet in an adult Evers SE, Orchard JW, Haddad RG. 1985. Bone
Medieval (12th–14th century) population from density in postmenopausal North American
Mediterranean France. Archives of Oral Biology Indian and Caucasian females. Human Biology
54:287–297. 57:719–726.
Eshed V, Gopher A, Galili E, Hershkovitz I. 2004. Evinger S, Bernert Z, Fothi E, et al. 2011. New
Musculoskeletal stress markers in Natufian skeletal tuberculosis cases in past populations
hunter-gatherers and Neolithic farmers in the from western Hungary (Transdanubia). Homo
Levant: the upper limb. American Journal of 62:165–183.
Physical Anthropology 123:303–315. Eyre-Brook AL. 1984. The periosteum: its function
Eshed V, Gopher A, Hershkovitz I. 2006. Tooth reassessed. Clinical Orthopaedics and Related
wear and dental pathology at the advent of Research 189:300–307.
agriculture: new evidence from the Levant. Ezzo JA. 1992. Dietary change and variability at
American Journal of Physical Anthropology Grasshopper Pueblo, Arizona. Journal of
130:145–159. Anthropological Archaeology 11:219–289.
Eshed V, Gopher A, Pinhasi R, Hershkovitz I. 2010. Ezzo JA. 1993. Human Adaptation at Grasshopper
Paleopathology and the origin of agriculture in Pueblo, Arizona: Social and Ecological
the Levant. American Journal of Physical Perspectives. International Monographs in
Anthropology 143:121–133. Prehistory, Archaeological Series, No. 4.
Espallargues M, Sampietro-Colom L, Estrada MD, Ezzo JA. 1994a. Putting the “chemistry” back
et al. 2001. Identifying bone mass-related risk into archaeological chemistry analysis:
factors for fracture to guide bone densitometry modeling potential paleodietary indicators.
measurements: a systematic review of the Journal of Anthropological Archaeology
literature. Osteoporosis International 13:1–34.
12:811–822. Ezzo JA. 1994b. Zinc as a paleodietary indicator:
Estabrook VH, Frayer DW. 2014. Trauma in the an issue of theoretical validity in bone-
Krapina Neandertals: violence in the Middle chemistry analysis. American Antiquity
Paleolithic? In: Knüsel C, Smith MJ, editors. 59:606–621.
The Routledge Handbook of the Bioarchaeology Ezzo JA, Larsen CS, Burton JH. 1995. Elemental
of Human Conflict. London, UK: Routledge; signatures of human diets from the Georgia
p. 67–89. Bight. American Journal of Physical
Evans J, Stoodley N, Chenery C. 2006. A strontium Anthropology 98:471–481.
and oxygen isotope assessment of a possible Fahlström G. 1981. The glenohumeral joint in
fourth century immigrant population in a man: an anatomic-experimental and archaeo-
Hampshire cemetery, southern England. osteological study on joint function. Ossa,
Journal of Archaeological Science 33: Supplement 1.
265–272. Fairgrieve SI, Molto JE. 2000. Cribra orbitalia in
Evans MW. 1944. Congenital dental defects in two temporally distinct population samples
infants subsequent to maternal rubella during from the Dakhleh Oasis, Egypt. American
pregnancy. Medical Journal of Australia Journal of Physical Anthropology
2:225–228. 111:319–331.
Eveleth PB, Tanner JM. 1990. Worldwide Fairservis WAJ. 1971. The Roots of Ancient India:
Variation in Human Growth, Second Edition. The Archaeology of Early Indian Civilization.
Cambridge, UK: Cambridge University Press. New York, NY: MacMillan.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
470 References

Fallon MD. 1988. Bone histomorphology. In: editors. Osteoarthritis, Second Edition. New
Resnick D, Niwayama G, editors. Diagnosis of York, NY: Oxford University Press; p. 9–16.
Bone and Joint Disorders, Second Edition. Felson DT, Anderson JJ, Naimark A, Walker AM,
Philadelphia, PA: Lea & Febiger; p. 1975–1997. Meenan RF. 1988. Obesity and knee
Farmer P. 1996. On suffering and structural osteoarthritis: the Framingham study. Annals
violence: a view from below. Daedalus of Internal Medicine 109:18–24.
125:261–283. Ferguson RB. 1990. Blood of the Leviathan:
Farmer P. 2003. Pathologies of Power: Health, Western contact and warfare in Amazonia.
Human Rights, and the New War on the Poor. American Ethnologist 17:237–257.
Berkeley, CA: University of California Ferguson RB. 1995. Yanomami Warfare:
Press. A Political History. Sante Fe, NM: School of
Farnsworth KB, Asch DL. 1986. Early Woodland American Research Press.
chronology, artifact styles, and settlement Ferguson RB, Whitehead N. 1992. War in the
distribution in the lower Illinois Valley region. Tribal Zone. Santa Fe, NM: School of American
In: Farnsworth KB, Emerson TE, editors. Early Research.
Woodland Archaeology. Center for American Ferrozo G, Morandi A, Bounaguro V, Jaafar J,
Archeology, Seminars in Archeology, No. 2; Draghi F. 1990. Significance of Harris lines in
p. 326–357. fractures of the lower limbs. Radiologia Medica
Farnsworth KB, Emerson TE, Glenn RM. 1991. 80:638–644.
Patterns of Late Woodland/Mississippian Fibiger L. 2014. Misplaced childhood?
interaction in the lower Illinois Valley Interpersonal violence and children in
drainage: a view from Starr Village. In: Neolithic Europe. In: Knüsel C, Smith MJ,
Emerson TE, Lewis RB, editors. Cahokia and the editors. The Routledge Handbook of the
Hinterlands. Urbana, IL: University of Illinois Bioarchaeology of Human Conflict. London,
Press; p. 83–118. UK: Routledge; p. 127–145.
Farnsworth P, Brady JE, DeNiro MJ, MacNeish RS. Fibiger L, Ahlstrom T, Bennike P, Schulting RJ.
1985. A re-evaluation of the isotopic and 2013. Patterns of violence-related skull trauma
archaeological reconstruction of diet in the in Neolithic southern Scandinavia. American
Tehuacan Valley. American Antiquity Journal of Physical Anthropology
50:102–116. 150:190–202.
Faulkner DK. 1986. The mass burial: an Field JS, Cochrane EE, Greenlee DM. 2009. Dietary
entomological perspective. In: Donnan CB, change in Fijian prehistory: isotopic analyses
Cook GA, editors. The Pacatnamu Papers, of human and animal skeletal material. Journal
Volume 1. Los Angeles, CA: Museum of of Archaeological Science 36:1547–1556.
Cultural History, University of California; Fienup-Riordan A. 1994. Eskimo war and peace.
p. 145–150. In: Fitzhugh WW, Chaussonnet V, editors.
Fearn ML, Liu K-B. 1995. Maize pollen of 3500 BP Anthropology of the North Pacific Rim.
from southern Alabama. American Antiquity Washington, DC: Smithsonian Institution
60:109–117. Press; p. 321–335.
Fejerskov O, Thylstrup A. 1986. Dental enamel. In: Fildes V. 1995. The culture and biology of
Mjör IA, Fejerskov O, editors. Human Oral breastfeeding: an historical overview of
Embryology. Copenhagen: Munksgaard western Europe. In: Stuart-Macadam P,
International Publishers; p. 50–89. Dettwyler KA, editors. Breastfeeding:
Felson DT. 2000. Osteoarthritis: the disease and its Biocultural Perspectives. New York, NY: Aldine
prevalence and impact. Annals of Internal de Gruyter; p. 76–101.
Medicine 133:635–636. Filon D, Faerman M, Smith P, Oppenheim A. 1995.
Felson DT. 2003. Epidemiology of osteoarthritis. Sequence analysis reveals a B-thalassaemia
In: Brandt KD, Doherty M, Lohmander LS, mutation in the DNA of skeletal remains from

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 471

the archaeological site of Akhziv, Israel. Nature Human Skeleton, Second Edition. Hoboken, NJ:
Genetics 9:365–368. Wiley-Liss; p. 237–263.
Fine D, Craig GT. 1981. Buccal surface wear of Fitzhugh W. 1994. Foreword. In: Bray TL, Killion
human premolar and molar teeth: a potential TW, editors. Reckoning with the Dead: The
indicator of dietary and social differentiation. Larsen Bay Repatriation and the Smithsonian
Journal of Human Evolution 10:335–344. Institution. Washington, DC: Smithsonian
Fink TM. 1985. Tuberculosis and anemia in a Institution Press; p. vii–vix.
Pueblo II–III (ca. AD 900–1300) Anasazi child Fitzhugh W, Ward E. 2000. Vikings: the North
from New Mexico. In: Merbs CF, Miller RJ, Atlantic Saga. Washington, DC: Smithsonian
editors. Health and Disease in the Prehistoric Institution Press.
Southwest. Arizona State University Fitzpatrick SM, Ross AH, editors. 2011. Island
Anthropological Research Papers, No. 34; Shores, Distant Pasts: Archaeological and
p. 359–379. Biological Approaches to the Pre-Columbian
Finnegan M. 1978. Non-metric variation of the Settlement of the Caribbean. Gainesville, FL:
infracranial skeleton. Journal of Anatomy University Press of Florida.
125:23–37. Fletcher HA, Donoghue HD, Holton J, Pap I,
Finucane BC. 2007. Mummies, maize, and Spigelman M. 2003. Widespread occurrence of
manure: multi-tissue stable isotope analysis of Mycobacterium tuberculosis DNA from 18th–
late prehistoric human remains from the 19th century Hungarians. American Journal of
Ayacucho Valley, Peru. Journal of Physical Anthropology 102:144–152.
Archaeological Science 34:2115–2124. Flores RH, Hochberg MC. 2003. Definition and
Fiorato V, Boylston A, Knüsel C, editors. 2000. classification of osteoarthritis. In: Brandt KD,
Blood Red Roses: The Archaeology of a Mass Doherty M, Lohmander LS, editors.
Grave from the Battle of Towton AD 1461. Osteoarthritis, Second Edition. New York, NY:
Oxford, UK: Oxbow Books. Oxford University Press; p. 1–8.
Fischer A, Olsen J, Richards M, et al. 2007. Floud R. 1994. The heights of Europeans since
Coast–inland mobility and diet in the Danish 1750: a new source for European economic
Mesolithic and Neolithic: evidence from history. In: Komlos J, editors. Stature, Living
stable isotope values of humans and dogs. Standards, and Economic Development: Essays
Journal of Archaeological Science in Anthropometric History. Chicago, IL:
34:2125–2150. University of Chicago Press; p. 9–24.
Fischman J. 1996. California social climbers: low Floud R, Wachter K, Gregory A. 1990. Height,
water prompts high status. Science Health and History: Nutritional Status in the
272:811–812. United Kingdom, 1750–1980. Cambridge, UK:
FitzGerald C, Saunders S, Bondioli L, Macchiarelli Cambridge University Press.
R. 2006. Health of infants in an Imperial Fogel M, Tuross N, Owsley D. 1989. Nitrogen
Roman skeletal sample: perspective from Isotope Tracers of Human Lactation in Modern
dental microstructure. American Journal of and Archaeological Populations. Carnegie
Physical Anthropology 130:179–189. Institution of Washington, Geophysical
FitzGerald CM. 1998. Do enamel microstructures Laboratory, Annual Report of the Director,
have regular time dependency? Conclusions 1988–89; p. 111–116.
from the literature and large-scale study. Fogel ML, Tuross N, Johnson BJ, Miller GH. 1997.
Journal of Human Evolution 35:371–386. Biogeochemical record of ancient humans.
FitzGerald CM, Rose JC. 2008. Reading between Organic Geochemistry 27:275–287.
the lines: dental development and subadult age Fogel RW, Engerman SL, Floud R, et al. 1983.
assessment using the microstructural growth Secular changes in American and British
markers of teeth. In: Katzenberg MA, Saunders stature and nutrition. Journal of
SR, editors. Biological Anthropology of the Interdisciplinary History 14:445–481.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
472 References

Food and Agriculture Organization (FAO). 1953. Bolivia. American Journal of Physical
Maize and Maize Diets: A Nutritional Survey. Anthropology 126:343–351.
Rome, Italy: FAO. Fowke G. 1902. Archaeological History of Ohio.
Food and Agriculture Organization (FAO). 1970. Columbus, OH: Ohio State Archaeological and
Amino-Acid Content on Foods and Biological Historical Society.
Data on Proteins. Rome, Italy: FAO. Fowler CS. 1992. In the Shadow of Fox Peak: An
Forgey K. 2011. Nasca trophy head origins and Ethnography of the Cattail-Eater Northern
ancient DNA. In: Bonogofsky M, editor. The Paiute People of Stillwater Marsh. Washington,
Bioarchaeology of the Human Head: DC: Department of Interior, Fish, and Wildlife
Decapitation, Decoration, and Deformation. Service.
Gainesville, FL: University Press of Florida; Fowler ML, Rose J, Vander Leest B, Ahler SR.
p. 286–306. 1999. The Mound 72 Area: Dedicated and
Formicola V. 1986–1987. The dentition of the Sacred Space in Early Cahokia. Illinois State
Neolithic sample from western Liguria (Italy). Museum, Reports of Investigations, No. 54.
Ossa 13:97–107. Fowler WR Jr. 1984. Late Preclassic mortuary
Formicola V, Milanesi Q, Scarsini C. 1987. patterns and evidence for human sacrifice at
Evidence of spinal tuberculosis at the Chalchuapa, El Salvador. American Antiquity
beginning of the fourth millennium B.C. from 49:603–618.
Arene Candide Cave (Liguria, Italy). American Fox CL. 1996. Physical anthropological aspects of
Journal of Physical Anthropology 72:1–6. the Mesolithic–Neolithic transition in the
Fornaciari G. 2008. Food and disease at the Iberian peninsula. Current Anthropology
Renaissance courts of Naples and Florence: a 37:689–695.
paleonutritional study. Appetite 51:10–14. Frake SE, Goose DH. 1977. A comparison between
Fornaciari G, Giuffra V, Ferroglio E, Bianucci R. Medieval and modern British mandibles.
2010a. Malaria was “the killer” of Francesco I Archives of Oral Biology 22:55–57.
de’ Medici. American Journal of Medicine Francalacci P, Borgognini Tarli S. 1988.
1232:568–569. Multielement analysis of trace elements and
Fornaciari G, Giuffra V, Ferroglio E, Gino S, preliminary results on stable isotopes in two
Bianucci R. 2010b. Plasmodium falciparum Italian prehistoric sites: methodological
immunodetection in bone remains of members aspects. In: Grupe G, Hermann B, editors. Trace
of the Renaissance Medici family (Florence, Elements in Environmental History. Berlin,
Italy, sixteenth century). Transactions of the Germany: Springer-Verlag; p. 41–52.
Royal Society of Tropical Medicine and Hygiene Franceschi MG, Tofanelli S, Paoli G. 1994.
104:583–587. Variance of ABO gene frequencies in Lower
Fornaciari G, Mallegni F, Bertini D, Nuti V. 1981. Nubia through the last millennium. Homo
Cribra orbitalia, and elemental bone iron, in the 45:51–62.
Punics of Carthage. Ossa 8:63–77. Frankenberg SR, Konigsberg LW. 2006. A brief
Forsberg CM, Eliasson S, Westergren H. 1991. Face history of paleodemography from Hooton to
height and tooth eruption in adults – a 20-year hazards analysis. In: Buikstra JE, Beck LA,
follow-up investigation. European Journal of editors. Bioarchaeology: The Contextual
Orthodontics 13:249–254. Analysis of Human Remains. Burlington, VT:
Forsberg K, Nilsson BE. 1992. Coxarthrosis on the Academic Press; p. 227–254.
island of Gotland: increased prevalence in a Frayer DW. 1977. Metric dental change in the
rural population. Acta Orthopaedica European Upper Paleolithic and Mesolithic.
Scandinavica 63:1–3. American Journal of Physical Anthropology
Foster Z, Byron E, Reyes-Garcia V, et al. 2005. 46:109–120.
Physical growth and nutritional status of Frayer DW. 1978. Evolution of the Dentition in
Tsimane’ Amerindian children of lowland Upper Paleolithic and Mesolithic Europe.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 473

University of Kansas Publications in Froment A. 1994. Épidémiologie des


Anthropology, No. 10. tréponématoses endémiques Africaines de
Frayer DW. 1984. Tooth size, oral pathology and savane et de forêt. In: Dutour O, Pálfi G, Bérato
class distinctions: evidence from the J, Brun J-P, editors. L’Origine de la Syphilis en
Hungarian Middle Ages. Anthropologiai Europe: Avant ou Après 1493? Paris, France:
Közlemenyek 28:47–54. Éditions Errance; p. 41–47.
Frayer DW. 1988a. Caries and oral pathologies at Frost HM. 1966. Morphometry of bone in
the Mesolithic sites of Muge: Cabeço da Arruda palaeopathology. In: Jarcho S, editor. Human
and Moita do Sebastião. Trabajo Antropologia Palaeopathology. New Haven, CT: Yale
e Etnologia 27:9–25. University Press; p. 131–150.
Frayer DW. 1988b. Sex differences in tooth wear Frost HM. 2003. On changing views about age-
at Skateholm. Mesolithic Miscellany 9:11–12. related bone loss. In Agarwal SC, Stout SD,
Frayer DW. 1997. Ofnet: evidence for a Mesolithic editors. Bone Loss and Osteoporosis: An
massacre. In: Martin DL, Frayer DW, editors. Anthropological Perspective. New York, NY:
Troubled Times: Violence and Warfare in the Kluwer Academic Plenum; p. 19–31.
Past. Amsterdam, the Netherlands: Gordon & Fry B. 2006. Stable Isotope Ecology. New York,
Breach; p. 181–216. NY: Springer.
Frayer DW, Lozano M, Bermúdez de Castro JM, Fujita H. 1995. Geographical and chronological
et al. 2012. More than 500,000 years of right- differences in dental caries in the Neolithic
handedness in Europeans. Laterality 17:51–69. Jomon period of Japan. Anthropological
Fredrickson BE, Baker D, McHolick WJ, Yuan HA, Science 103:23–37.
Lubicky JP. 1984. The natural history of Fujita H, Ogura M. 2009. Degree of dental attrition
spondylolysis and spondylolisthesis. Journal of with sex and aging among Jomon and Edo
Bone and Joint Surgery 66-A:699–707. people in Japan. Journal of Oral Biosciences
Fresia AE, Ruff CB, Larsen CS. 1990. Temporal 51:165–171.
decline in bilateral asymmetry of the upper Fukase H, Suwa G. 2008. Growth-related changes
limb on the Georgia coast. In: Larsen CS, editor. in prehistoric Jomon and modern Japanese
The Archaeology of Mission Santa Catalina de mandibles with emphasis on cortical bone
Guale: 2. Biocultural Interpretations of a distribution. American Journal of Physical
Population in Transition. Anthropological Anthropology 136:441–454.
Papers of the American Museum of Natural Fuller BT, Fuller JL, Harris DA, Hedges REM. 2006.
History, No. 68; p. 121–132. Detection of breastfeeding and weaning in
Fricke HC, O’Neil JR, Lynnerup N. 1995. Oxygen modern human infants with carbon and
isotope composition of human tooth enamel nitrogen stable isotope ratios. American Journal
from Medieval Greenland: linking climate and of Physical Anthropology 129:279–293.
society. Geology 23:869–872. Fuller BT, Márquez-Grant N, Richards MP. 2010.
Friedmann LW. 1973 Amputation in pre- Investigation of diachronic dietary patterns on
Columbian America. Archives of Physical the islands of Ibiza and Formentera, Spain:
Medicine and Rehabilitation 54:323–325. evidence from carbon and nitrogen stable
Frisancho AR, Garn SM, Ascoli W. 1970. isotope ratio analysis. American Journal of
Subperiosteal and endosteal bone apposition Physical Anthropology 143:512–522.
during adolescence. Human Biology Fuller BT, Richards MP, Mays SA. 2003. Stable
42:639–664. carbon and nitrogen isotope variations in tooth
Frisancho AR, Newman MT, Baker P. 1970. dentine serial sections from Wharram Percy.
Differences in stature and cortical thickness Journal of Archaeological Science
among highland Quechua Indian boys. 30:1673–1684.
American Journal of Clinical Nutrition Furst CM, Hansen FCC. 1915. Crania
23:382–385. Groenlandica: A Description of Greenland

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
474 References

Eskimo Crania. Copenhagen, Denmark: A.F. Garn SM, Cole PE, Smith BH. 1979. The effect of
Höst and Son. sample size on crown size asymmetry. Journal
Gadd CJ. 1932. Seals of the ancient Indian style of Dental Research 58:2012.
found at Ur. Proceedings of the British Garn SM, Lewis AB, Kerewsky S. 1965. Genetic,
Academy 17:191–210. nutritional, and maturational correlates of
Gage T. 2010. Demographic estimation: indirect dental development. Journal of Dental
techniques for anthropological populations. In: Research 44:228–242.
Larsen CS, editor. A Companion to Biological Garn SM, Lewis AB, Polacheck DL. 1960.
Anthropology. Malden, MA: Wiley-Blackwell; Interrelations in dental development.
p. 179–193. I. Interrelationships within the dentition.
Gage TB, DeWitte S. 2009. What do we know Journal of Dental Research 39:1040–1055.
about the agricultural demographic transition? Garn SM, Osborne RH, Alvesalo L, Horowitz SL.
Current Anthropology 50:649–655. 1980. Maternal and gestational influences on
Gagnon CM, Wiesen C. 2013. Using general deciduous and permanent tooth size. Journal of
estimating equasions to analyze oral health in Dental Research 59:142–143.
the Moche Valley of Peru. International Garn SM, Osborne RH, McCabe KD. 1979 The
Journal of Osteoarchaeology, 23:557–572. effect of prenatal factors on crown dimensions.
Gaither CM, Murphy MS. 2012. Consequences of American Journal of Physical Anthropology
conquest? The analysis and interpretation of 51:665–678.
subadult trauma at Puruchuco-Huaquerones, Garn SM, Rohmann CG, Behar M, Viteri F,
Peru. Journal of Archaeological Science Guzman MA. 1964. Compact bone deficiency
39:467–478. in protein-calorie malnutrition. Science
Galloway A. 1999. Broken Bones: Anthropological 145:144–145.
Analysis of Blunt Force Trauma. Springfield, Garn SM, Rohmann CG, Wagner B, Ascoli W.
IL: Charles C Thomas. 1967. Continuing bone growth throughout life:
Gardner JC, Smith FH. 2006. The paleopathology a general phenomenon. American Journal of
of the Krapina Neandertals. Periodicum Physical Anthropology 26:313–318.
Biologorum 108:471. Garn SM, Rohmann CG, Wagner E, Davila GH.
Garn SM. 1970. The Earlier Gain and Later Loss of 1970. Dynamics of change at the endosteal
Cortical Bone. Springfield, IL: Charles C. surface of tubular bones. In: Whedon D,
Thomas. Cameron JR, editors. Progress in Methods of
Garn SM. 1976. The insides of the bones and the Bone Mineral Measurement. Washington, DC:
mass of humanity. Annals of the XLI US Government Printing Office; p. 430–453.
International Congress of Americanists Garn SM, Schwager PM. 1967. Age dynamics of
3:453–457. persistent transverse lines in the tibia.
Garn SM. 1985. Comment. Current Anthropology American Journal of Physical Anthropology
26:350–351. 27:375–378.
Garn SM. 1989. Directions of aging. In: Carlson Garn SM, Silverman FN, Hertzog KP, Rohmann
DS, editor. Orthodontics in an Aging Society. VM. 1968. Lines and bands of increased
University of Michigan, Center for Human density: their implication to growth and
Growth and Development, Craniofacial Growth development. Medical Radiography and
Series Monograph, No. 22; p. 33–39. Photography 44:58–89.
Garn SM, Baby RS. 1969. Bilateral symmetry in Garn SM, Sullivan TV, Decker SA, Larkin FA,
finer lines of increased density. American Hawthorne VM. 1992. Continuing bone
Journal of Physical Anthropology 31:89–92. expansion and increasing bone loss over a
Garn SM, Burdi AR. 1971. Prenatal ordering and two-decade period in men and women from a
postnatal sequence in dental development. total community sample. American Journal of
Journal of Dental Research 50:1407–1414. Human Biology 4:57–67.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 475

Garner C. 1991. Kuhlman demography and disease Giblin JI. 2009. Strontium isotope analysis of
in comparative perspective. In: Atwell KA, Neolithic and Copper Age populations on the
Conner MD, editors. The Kuhlman Mound Great Hungarian Plain. Journal of
Group and Late Woodland Mortuary Archaeological Science 36:491–497.
Behavior in the Mississippi River Valley of Gibson KR. 1984. Comment. Current Anthropology
West-Central Illinois. Center for American 25:320–321.
Archaeology, Research Series, No. 9; Gilbert RI Jr. 1975. Trace Element Analysis of
p. 180–207. Three Skeletal Amerindian Populations at
Garofalo EM. 2013. Environmental and Genetic Dickson Mounds. PhD Dissertation, University
Effects on Growth of the Human Skeleton – A of Massachusetts, Amherst, MA.
Bioarchaeological Investigation. PhD Gilbert RI Jr. 1977. Applications of trace element
dissertation, Johns Hopkins University, research to problems in archeology. In: Blakely
Baltimore, MD. RL, editor. Biocultural Adaptation in
Garrelt C, Wiechmann I. 2003. Detection of Prehistoric America. Athens, GA: University of
Yersinia pestis DNA in early and late Medieval Georgia Press; p. 85–100.
Bavarian burials. In: Grupe G, Peters J, editors. Gilbert RI Jr. 1985. Stress, paleonutrition, and
Deciphering Ancient Bones: The Research trace elements. In: Gilbert RI, Mielke JH,
Potential of Bioarchaeological Collections. editors. The Analysis of Prehistoric Diets.
Rahden, Germany: Verlag Marie Leidorf Orlando, FL: Academic Press; p. 339–358.
GmbH; p. 247–254. Gill GW, Gilbert BM. 1990. Race identification
Gasperetti MA, Sheridan SG. 2013. Cry havoc: from the midfacial skeleton: American
interpersonal violence at Early Bronze Age Bab Blacks and Whites. In: Gill GW, Rhine S,
edh-Dhra’. American Anthropologist editors. Skeletal Attribution of Race.
115:388–410. Albuquerque, NM: University of New Mexico,
Gaur R, Kumar P. 2012. Effect of undernutrition Maxwell Museum of Anthropology;
on deciduous tooth emergence among Rajput p. 47–53.
children of Shimla District of Himachal Gill GW, Owsley DW. 1993. Human osteology of
Pradesh, India. American Journal of Physical Rapanui. In: Fischer SR, editor. Easter Island
Anthropology 148:54–61. Studies: Contributions to the History of
Geber J, Murphy E. 2012. Scurvy in the Great Irish Rapanui in Memory of William T. Mulloy.
famine: evidence of vitamin C deficiency from Oxbow Monograph, No. 32; p. 56–62.
a mid-19th century population. American Gilmore CC, Grote MN. 2012. Estimating age from
Journal of Physical Anthropology adult occlusal wear: a modification of the Miles
148:512–524. method. American Journal of Physical
Geddes DAM. 1994. Diet patterns and caries. Anthropology 149:181–192.
Advances in Dental Research 8:221–224. Gimbutas M. 1982. The Goddesses and Gods of Old
Gejvall N-G. 1960. Westerhus. Lund, Sweden: Europe. London, UK: Thames & Hudson.
Ohlssons. Ginter JK. 2011. Using a bioarchaeological
Gelber RH. 2008. Leprosy (Hansen’s disease). In: approach to explore subsistence transitions in
Faunci AS, Braunwald E, Kasper DL, et al. the eastern Cape, South Africa during the
editors. Harrison’s Principles of Internal mid- to late Holocene. In: Pinhasi R, Stock JT,
Medicine, Seventeenth Edition. New York, NY: editors. Human Bioarchaeology of the
McGraw-Hill; p. 1021–1027. Transition to Agriculture. Chichester, UK:
Gerry JP, Krueger HW. 1997. Regional diversity in Wiley-Blackwell; p. 107–149.
Classic Maya diets. In: Whittington SL, Reed Girotti M, Doro Garetto T. 1999. Studio
DM, editors. Bones of the Maya: Studies of odontologico sulla populazione. In: Negro
Ancient Skeletons. Washington, DC: Ponzi Mancini MM, editor. San Michele di
Smithsonian Institution Press; p. 196–207. Trino (VC, Dal Villaggio Romano al Castello

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
476 References

Medievale, Volume 2). Florence, Italy: editor. Bioarchaeological Studies of Life in the
All’insegna del Giglio; p. 732–738. Age of Agriculture. Tuscaloosa, AL: University
Giuffra V, Baricco LP, Subbrizio M, Fornaciari G. of Alabama Press; p. 195–218.
2013. Weapon-related cranial lesions from Gold DL. 2004. The Bioarchaeology of Virginia
Medieval and Renaissance Turin, Italy. Burial Mounds. Tuscaloosa, AL: University of
International Journal of Osteoarchaeology Alabama Press.
DOI:10.1002/oa.2334. Goldschmidt W, Foster G, Essene F. 1939. War
Giuffra V, Vitiello A, Caramella D, et al. 2013. stories from two enemy tribes. Journal of
Rickets in a high social class of Renaissance American Folklore 52:141–154.
Italy: the Medici children. International Goldstein L. 2006. Mortuary analysis and
Journal of Osteoarchaeology DOI:10.1002/ bioarchaeology. In: Buikstra JE, Beck LA,
oa.2324. editors. Bioarchaeology: The Contextual
Gladykowska-Rzeczycka JJ. 1999. Tuberculosis in Analysis of Human Remains. Amsterdam, the
the past and present in Poland. In: Pálfi G, Netherlands: Elsevier; p. 375–387.
Dutour O, Deak J, Hutas I, editors. Tuberculosis: Goldstein MS. 1936. Changes in dimensions and
Past and Present. Budapest/Szeged, Hungary: form of the face and head with age. American
Golden Book Publishers and Tuberculosis Journal of Physical Anthropology 22:27–89.
Foundation; p. 561–573. Goldstein MS. 1953. Some vital statistics based on
Glass RI, Urrutia JJ, Sibony S, et al. 1977. skeletal material. Human Biology 25:3–12.
Earthquake injuries related to housing in a Goldstein P. 1992. Tiwanaku expansion and state
Guatemalan village. Science 197:638–643. expansion. Latin American Antiquity 4:22–47.
Glassow MA. 1996. Purisimeño Chumash González-José R, Neves W, Lahr MM, et al. 2005.
Prehistory: Maritime Adaptations along the Late Pleistocene/Holocene craniofacial
Southern California Coast. Fort Worth, TX: morphology in Mesoamerican Paleoindians:
Harcourt Brace College Publishers. implications for the peopling of the New
Glencross B, Agarwal SC. 2011. An investigation World. American Journal of Physical
of cortical bone loss and fracture patterns in Anthropology 128:772–780.
the Neolithic community of Çatalhöyük, González-José R, Ramírez-Rozzi F, Sardi M, et al.
Turkey using metacarpal radiogrammetry. 2005. Functional-cranial approach to the
Journal of Archaeological Science 38:513–521. influence of economic strategy on skull
Glencross BA. 2011. Skeletal injury across the life morphology. American Journal of Physical
course. In: Agarwal SC, Glencross BA, editors. Anthropology 128:757–771.
Social Bioarchaeology. Chichester, UK: Wiley- Goodman AH. 1989. Dental enamel hypoplasias in
Blackwell; p. 392–411. prehistoric populations. Advances in Dental
Glob PV. 1971. The Bog People: Iron-Age Man Research 3:265–271.
Preserved. New York, NY: Ballantine Books. Goodman AH. 1991. Health, adaptation, and
Gluckman PD, Hanson MA, Cooper C, Thornburg maladaptation in past societies. In: Bush H,
KL. 2008. Effect of in utero and early-life Zvelebil M, editors. Health in Past Societies:
conditions on adult health and disease. New Biocultural Interpretations of Human Skeletal
England Journal of Medicine 359:61–73. Remains in Archaeological Contexts. British
Godde K, Hens SM. 2012. Age-at-death estimation Archaeological Reports, International Series,
in an Italian historical sample: a test of the No. 567; p. 31–38.
Suchey-Brooks and transition analysis Goodman AH. 1993. On the interpretation of
methods. American Journal of Physical health from skeletal remains. Current
Anthropology 149:259–265. Anthropology 34:281–288.
Gold DL. 2000. “Utmost confusion” reconsidered: Goodman AH. 1994. Cartesian reductionism and
bioarchaeology and secondary burial in late vulgar adaptationism: issues in the
prehistoric internal Virginia. In: Lambert PM, interpretation of nutritional status in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 477

prehistory. In: Sobolik KD, editor. Goodman AH, Clark GA. 1981. Harris lines as
Paleonutrition: The Diet and Health of indicators of stress in prehistoric Illinois
Prehistoric Americans. Southern Illinois populations. In: Martin DL, Bumsted MP,
University at Carbondale, Center for editors. Biocultural Adaptation:
Archaeological Investigations, Occasional Comprehensive Approaches to Skeletal
Paper, No. 22; p. 163–177. Analysis. University of Massachusetts,
Goodman AH. 1998. The biological consequences Department of Anthropology, Research
of inequality in antiquity. In Goodman AH, Reports, No. 20; p. 35–46.
Leatherman TL, editors. Building a New Goodman AH, Jones J, Reid J, et al. 2009. Isotopic
Biocultural Synthesis: Political-Economic and elemental chemistry of teeth: implications
Perspectives on Human Biology. Ann Arbor, for places of birth, forced migration patterns,
MI: University of Michigan Press; nutritional status, and pollution. In: Blakey
p. 147–169. ML, Rankin-Hill LM, editors. The Skeletal
Goodman AH. 2013. Bringing culture into human Biology of the New York African Burial
biology and biology back into anthropology. Ground, Part I. Washington, DC: Howard
American Anthropologist 115:359–373. University Press; p. 95–118.
Goodman AH, Allen LH, Hernandez GP, et al. Goodman AH, Lallo J, Armelagos GJ, Rose JC.
1987. Prevalence and age at development of 1984. Health changes at Dickson Mounds,
enamel hypoplasias in Mexican children. Illinois (A.D. 950–1300). In: Cohen MN,
American Journal of Physical Anthropology Armelagos GJ, editors. Paleopathology at the
72:7–19. Origins of Agriculture. Orlando, FL: Academic
Goodman AH, Armelagos GJ. 1985a. The Press; p. 271–305.
chronological distribution of enamel Goodman AH, Martin DL. 2002. Reconstructing
hypoplasia in human permanent incisor and health profiles from skeletal remains. In:
canine teeth. Archives of Oral Biology Steckel RH, Rose JC, editors. The Backbone of
30:503–507. History: Health and Nutrition in the Western
Goodman AH, Armelagos GJ. 1985b. Factors Hemisphere. New York, NY: Cambridge
affecting the distribution of enamel University Press; p. 11–60.
hypoplasias within the human permanent Goodman AH, Martin DL, Armelagos GJ, Clark C.
dentition. American Journal of Physical 1984. Indications of Stress from Bone and
Anthropology 68:479–493. Teeth. In: Cohen MN, Armelagos GJ, editors.
Goodman AH, Armelagos GJ. 1988. Childhood Paleopathology at the Origins of Agriculture.
stress, cultural buffering, and decreased Orlando, FL: Academic Press; p. 13–50.
longevity in a prehistoric population. American Goodman AH, Martinez C, Chavez A. 1991.
Anthropologist 90:936–944. Nutritional supplementation and the
Goodman AH, Armelagos GJ. 1989. Infant and development of linear enamel hypoplasias in
childhood morbidity and mortality risks in children from Tezonteopan, Mexico. American
archaeological populations. World Archaeology Journal of Clinical Nutrition 53:773–781.
21:225–243. Goodman AH, Pelto GH, Allen LH, Chavez A.
Goodman AH, Armelagos GJ, Rose JC. 1980. 1992. Socioeconomic and anthropometric
Enamel hypoplasias as indicators of stress in correlates of linear enamel hypoplasia in
three prehistoric populations from Illinois. children from Solis, Mexico. In: Goodman AH,
Human Biology 52:515–528. Capasso LL, editors. Recent Contributions to the
Goodman AH, Armelagos GJ, Rose JC. 1984. The Study of Enamel Developmental Defects.
chronological distribution of enamel Journal of Paleopathology, Monographic
hypoplasias from prehistoric Dickson Mounds Publications, No. 2; p. 373–380.
populations. American Journal of Physical Goodman AH, Rose JC. 1990. Assessment of
Anthropology 65:259–266. systemic physiological perturbations from

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
478 References

dental enamel hypoplasias and associated to detect diet. Journal of Dental Research
histological structures. Yearbook of Physical 63:1043–1046.
Anthropology 33:59–110. Gordon S, Mwandumba H. 2008. Respiratory
Goodman AH, Rose JC. 1991. Dental enamel tuberculosis. In: Davis PDO, Barnes PF, Gordon
hypoplasias as indicators of nutritional status. SB, editors. Clinical Tuberculosis. Oxford, UK:
In Kelley MA, Larsen CS, editors. Advances in Hodder Arnold; p. 145–162.
Dental Anthropology. New York, NY: Wiley- Gorsky M, Bukai A, Shohat M. 1998. Genetic
Liss; p. 279–293. influence on the prevalence of torus palatinus.
Goodman AH, Thomas RB, Swedlund AC, American Journal of Medical Genetics
Armelagos GJ. 1988. Biocultural perspectives 75:138–140.
on stress in prehistoric, historical, and Gosman JH, Hubbell ZR, Shaw CN, Ryan TM.
contemporary population research. Yearbook of 2013. Development of cortical bone geometry
Physical Anthropology 31:169–202. in the human femoral and tibial diaphysis.
Goodship AE, Lanyon LE, McFie H. 1979. Anatomical Record 296:774–787.
Functional adaptation of bone to increased Gosman JH, Ketcham RA. 2009. Patterns of
stress. Journal of Bone and Joint Surgery ontogeny of human trabecular bone from Sun
61-A:539–546. Watch village in the prehistoric Ohio Valley:
Goose DH. 1962. Reduction of palate size in general features of microarchitectural change.
modern populations. Archives of Oral Biology American Journal of Physical Anthropology
7:343–350. 138:318–332.
Goose DH. 1963. Dental measurement: an Gosman JH, Stout SD, Larsen CS. 2011. Skeletal
assessment of its value in anthropological biology over the life span: a view from the
studies. In: Brothwell DR, editor. Dental surfaces. Yearbook of Physical Anthropology
Anthropology. New York, NY: Pergamon Press; 54:86–98.
p. 125–148. Gotthard K, Nylin S. 1995. Adaptive plasticity and
Goose DH. 1967. Preliminary study of tooth size in plasticity as adaptation: a selective review of
families. Journal of Dental Research 46:959–962. plasticity in animal morphology and life
Goose DH. 1971. The inheritance of tooth size in history. Oikos 74:3–17.
British families. In: Dahlberg AA, editor. Dental Gould SJ. 1996. The Mismeasure of Man, Second
Morphology and Evolution. Chicago, IL: Edition. New York, NY: W.W. Norton
University of Chicago Press; p. 263–270. & Company.
Goose DH. 1972. Maxillary dental arch width in Gowland R, Knüsel C, editors. 2006. The Social
Chinese living in Liverpool. Archives of Oral Archaeology of Human Remains. Oxford, UK:
Biology 17:231–233. Oxbow.
Goose DH. 1981. Changes in the human face Gowland RL, Chamberlain AT. 2005. Detecting
breadth since the Medieval period in Britain. plague: palaeodemographic characterisation of
Archives of Oral Biology 26:757–758. a catastrophic death assemblage. Antiquity
Goose DH, Parry SE. 1974. Palate width in skulls 79:146–157.
from a recently excavated English Medieval Grauer AL. 1991. Life patterns of women from
site. Archives of Oral Biology 19:273–274. Medieval York. In: Walde D, Willows ND,
Gordon CC, Buikstra JE. 1981. Soil pH, bone editors. The Archaeology of Gender.
preservation, and sampling bias at mortuary Archaeological Association of the University of
sites. American Antiquity 48:566–571. Calgary, Proceedings of the 22nd Annual
Gordon DH. 1958. Prehistoric Background of Chacmool Conference; p. 407–413.
Indian Culture. Bombay, India: Mandhuri Grauer AL. 1993. Patterns of anemia and infection
Dhirajlal. from Medieval York, England. American
Gordon KD. 1984. Hominoid dental microwear: Journal of Physical Anthropology 91:
complications in the use of microwear analysis 203–213.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 479

Grauer AL, Roberts CA. 1996. Paleoepidemiology, Greene DL. 1984. Fluctuating dental asymmetry
healing, and possible treatment of trauma in and measurement error. American Journal of
the Medieval cemetery population of St. Helen- Physical Anthropology 65:283–289.
on-the-Walls, York, England. American Greene DL, Armelagos GJ. 1972. The Wadi Halfa
Journal of Physical Anthropology Mesolithic Population. University of
100:531–544. Massachusetts, Department of Anthropology,
Grauer AL, Stuart-Macadam P, editors. 1998. Sex Research Report, No. 11.
and Gender in Paleopathological Perspective. Greene DL, Ewing GH, Armelagos G. 1967.
Cambridge, UK: Cambridge University Press. Dentition of a Mesolithic population from
Gray JP, Wolfe LD. 1996. What accounts for Wadi Halfa, Sudan. American Journal of
population variation in height? In: Ember CR, Physical Anthropology 27:41–55.
Ember M, Peregrine PN, editors. Research Greene PE, Chisick MC, Aaron GRA. 1994.
Frontiers in Anthropology. Needham, MA: Comparison of oral health status and need for
Simon & Schuster; p. 103–120. dental care between abused neglected children
Grayson DK. 1990. Donner party deaths: a and non-abused non-neglected children.
demographic assessment. Journal of Pediatric Dentistry 16:41–45.
Anthropological Research 46:223–242. Greenlee DM. 2006. Dietary variation and
Grayson DK. 1993. Differential mortality and the prehistoric maize farming in the Middle Ohio
Donner Party disaster. Evolutionary Valley. In: Staller J, Tykot RH, Benz B, editors.
Anthropology 2:151–159. Histories of Maize. Amsterdam, The
Grayson DK. 1996. Human mortality in a natural Netherlands: Elsevier; p. 215–233.
disaster: the Willie Handcart Company. Journal Gregg JB, Gregg PS. 1987. Dry Bones: Dakota
of Anthropological Research 52:185–205. Territory Reflected. Sioux Falls, SD: Sioux
Grayson DK. 2011. The Great Basin: A Natural Printing.
Prehistory. Berkeley, CA: University of Gregoricka LA. 2013a. Geographic origins and
California Press. dietary transitions during the Bronze Age in
Green R, Suchey JM. 1976. The use of inverse sine the Oman peninsula. American Journal of
transformations in the analysis of non-metric Anthropology 152:353–369.
cranial data. American Journal of Physical Gregoricka LA. 2013b. Residential mobility and
Anthropology 45:61–68. social identity in the periphery: strontium
Green S, Green S, Armelagos GJ. 1974. Settlement isotope analysis of archaeological tooth enamel
and mortality of the Christian site of Meinarti. from southeastern Arabia. Journal of
Journal of Human Evolution 3:297–316. Archaeological Science 40:452–464.
Greenberg JH, Turner CG II, Zegura SL. 1986. The Gremillion KJ. 2002. The development and
settlement of the Americas: a comparison of dispersal of agricultural systems in the
the linguistic, dental, and genetic evidence. Woodland period Southeast. In: Anderson DG,
Current Anthropology 27:477–497. Mainfort RC Jr, editors. The Woodland
Greene DL. 1967. Dentition of Meroitic, X-Group Southeast. Tuscaloosa, AL: University of
and Christian Populations from Wadi Halfa, Alabama Press; p. 483–501.
Sudan. University of Utah Anthropological Greulich WW. 1976. Some secular changes in the
Papers, No. 85. growth of American-born and native Japanese
Greene DL. 1972. Dental anthropology of early children. American Journal of Physical
Egypt and Nubia. Journal of Human Evolution Anthropology 45:553–568.
1:315–324. Greulich WW, Thoms H. 1938. The dimensions
Greene DL. 1982. Discrete dental variations and of the pelvic inlet of 789 White females.
biological distances of Nubian populations. Anatomical Record 72:45–51.
American Journal of Physical Anthropology Griffin JB. 1943. The Fort Ancient Aspect: Its
58:75–79. Cultural and Chronological Position in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
480 References

Mississippi Valley Archaeology. Ann Arbor, Grupe G, Piepenbrink H. 1988. Trace element
MI: University of Michigan Press. contaminations in excavated bones by
Griffin JB. 1967. Eastern North American microorganisms. In: Grupe G, Herrmann B,
archaeology: a summary. Science editors. Trace Elements in Environmental
156:175–191. History. Berlin, Germany: Springer-Verlag;
Griffin MC, Lambert PM, Driscoll EM. 2001. p. 103–112.
Biological relationships and population history Grupe G, Piepenbrink H, Schoeninger MJ. 1989.
of native peoples in Spanish Florida and the Note on microbial influence on stable carbon
American Southeast. In: Larsen CS, editor. and nitrogen isotopes in bone. Applied
Bioarchaeology of Spanish Florida: The Impact Geochemistry 4:299.
of Colonialism. Gainesville, FL: University Grynpas MD. 2003. The role of bone quality on
Press of Florida; p. 226–273. bone loss and bone fragility. In: Agarwal SC,
Grin GI. 1956. Endemic syphilis and yaws. Bulletin Stout SD, editors. Bone Loss and
of the World Health Organization 15:959–973. Osteoporosis: An Anthropological Perspective.
Grine FE. 1986. Dental evidence for dietary New York, NY: Kluwer Academic Plenum;
differences in Australopithecus and p. 33–44.
Paranthropus: a quantitative analysis of Guagliardo MF. 1982a. Tooth crown size
permanent molar microwear. Journal of differences between age groups: a possible new
Human Evolution 15:783–822. indicator of stress in skeletal samples.
Grine FE, Kay RF. 1988. Early hominid diets from American Journal of Physical Anthropology
quantitative image analysis of dental 58:383–389.
microwear. Nature 333:765–768. Guagliardo MF. 1982b. Craniofacial Structure,
Grine FE, Sponheimer M, Ungar PS, Lee-Thorp J, Aging, and Dental Function: Their
Teaford MF. 2012. Dental microwear and stable Relationships in Adult Human Skeletal Series.
isotopes inform the paleoecology of extinct PhD Dissertation, University of Tennessee,
hominins. American Journal of Physical Knoxville, TN.
Anthropology 148:285–317. Gualandi PB. 1992. Food habits and dental disease
Grolleau-Raoux J-L, Crubezy E, Rouge D, in an Iron-Age population. Anthropologischer
Brugne J-F, Saunders SR. 1997. Harris lines: a Anzeiger 50:67–82.
study of age-associated bias in counting and Guatelli-Steinberg D. 2008. Using perikymata to
interpretation. American Journal of Physical estimate the duration of growth disruptions in
Anthropology 103:209–217. fossil hominin teeth: issues of methodology
Groves SE, Roberts CA, Lucy S, et al. 2013. and interpretation. In: Irish JD, Nelson GC,
Mobility histories of 7th–9th century AD editors. Technique and Application in Dental
people buried at early medieval Bamburgh, Anthropology. Cambridge, UK: Cambridge
Northumberland, England. American Journal University Press; p. 71–86.
of Physical Anthropology 151:462–476. Guatelli-Steinberg D, Larsen CS, Hutchinson DL.
Grüneberg H. 1952. Genetical studies on the 2004. Prevalence and the duration of
skeleton of the mouse, IV. Quasi-continuous linear enamel hypoplasia: a comparative
variations. Journal of Genetics 51:95–114. study of Neandertals and Inuit foragers.
Grupe G. 1995. Reconstructing migration in the Journal of Human Evolution 47:65–84.
Bell Beaker period by 87Sr/86Sr isotope ratios Guatelli-Steinberg D, Lukacs JR. 1999.
in teeth and bones. In: Radlandski RJ, Renz H, Interpreting sex differences in enamel
editors. Proceedings of the 10th International hypoplasia in human and non-human
Symposium on Dental Morphology. Berlin, primates: developmental, environmental, and
Germany: “M” Marketing Services, C. & cultural considerations. Yearbook of Physical
M. Bronne GbR; p. 338–342. Anthropology 42:73–126.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 481

Guatelli-Steinberg D, Reid DJ, Bishop TA, in 7500-year-old Neolithic sites. Science


Larsen CS. 2005. Anterior tooth growth 310:1016–1018.
periods in Neandertals were comparable to Haapasalo H, Sievanen H, Kannus P, et al. 1996
those of modern humans. Proceedings of the Dimensions and estimated mechanical
National Academy of Sciences characteristics of the humerus after long-term
102:14197–14202. loading. Journal of Bone and Mineral Research
Guatelli-Steinberg D, Sciulli PW, Edgar HHJ. 11:864–872.
2006. Dental fluctuating asymmetry in the Haas CJ, Zink A, Molnar E, et al. 2000. Molecular
Gullah: tests of hypotheses regarding evidence for different stages of tuberculosis in
developmental stability in deciduous vs. ancient bone samples from Hungary. American
permanent and male vs. female teeth. Journal of Physical Anthropology
American Journal of Physical Anthropology 113:293–304.
129:427–434. Haas CJ, Zink A, Szeimies U, Nerlich AG. 2002.
Guerrero E, Naji S, Bocquet-Appel J-P. 2008. The Molecular evidence of Mycobacterium leprae
signal of the Neolithic demographic transition in historic bone samples from south Germany.
in the Levant. In: Bocquet-Appel J-P, Bar- In: Roberts CA, Lewis ME, Manchester K,
Yosef O, editors. The Neolithic Demographic editors. The Past and Present of Leprosy:
Transition and Its Consequences. New York, Archaeological, Historical, Palaeopathological
NY: Springer; p. 57–80. and Clinical Approaches. British
Guliaev VI. 2003. Amazons in the Scythia: new Archaeological Reports, International Series,
finds at the middle Don, southern Russia. No. 1054; p. 287–292.
World Archaeology 35:112–125. Haas J, editor. 1990. The Anthropology of War.
Gunness-Hey M. 1980. The Koniag Eskimo New York, NY: Cambridge University Press.
presacral vertebral column: variations, Haas J, Creamer W. 1993. Stress and Warfare
anomalies, and pathologies. Ossa 7:99–118. among the Kayenta Anasazi of the Thirteenth
Gunness-Hey M. 1981. Sondylolysis in the Koniag Century A.D. Fieldiana Anthropology, No. 21.
Eskimo vertebral column. In: Martin DL, Haas JD. 2006. The effects of iron deficiency on
Bumsted MP, editors. Biocultural Adaptation: physical performance. In: Food and Nutrition
Comprehensive Approaches to Skeletal Board, Institute of Medicine, editors. Mineral
Analysis. University of Massachusetts Research Requirement for Military Personnel: Levels
Reports, No. 20; p. 16–23. Needed for Cognitive and Physical Performance
Gurdjian ES. 1973. Head Injury from Antiquity to during Garrison Training. Washington, DC:
the Present with Special Reference to National Academies Press; p. 451–461.
Penetrating Head Wounds. Springfield, IL: Habicht-Mauche J, Levendosky AA, Schoeninger
Charles C. Thomas. MJ. 1994. Antelope Creek Phase subsistence:
Gustafsson BE, Quensel C-E, Lanke LS, et al. 1954. the bone chemistry evidence. In: Owsley DW,
The effect of different levels of carbohydrate Jantz RL, editors. Skeletal Biology in the Great
intake on caries activity in 436 individuals Plains: Migration, Warfare, Health, and
observed for 5 years. Acta Odontologica Subsistence. Washington, DC: Smithsonian
Scandinavica 11:232–264. Institution Press; p. 291–304.
Haak W, Brandt G, de Jong HN, et al. 2008. Hackett CJ. 1936. Boomerang Leg and Yaws in
Ancient DNA, strontium isotopes, and Australian Aborigines. Royal Society of Tropical
osteological analyses shed light on social and Medicine & Hygiene, Monograph, No. 1.
kinship organization of the Later Stone Age. Hackett CJ. 1951. Bone Lesions of Yaws in
Proceedings of the National Academy of Uganda. Oxford, UK: Blackwell Scientific
Sciences 105:18226–18231. Publications.
Haak W, Forster P, Bramanti B, et al. 2005. Hackett CJ. 1976. Diagnostic Criteria of Syphilis,
Ancient DNA from the first European farmers Yaws and Treponarid (Treponematoses) and

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
482 References

Some Other Diseases in Dry Bones. Berlin, Handler JS, Aufderheide AC, Corruccini RS. 1986.
Germany: Springer-Verlag. Lead content and poisoning in Barbados slaves.
Hadler NM. 1977. Industrial rheumatology: Social Science History 10:399–425.
clinical investigations into the influence of Handler JS, Aufderheide AC, Corruccini RS,
patterns of usage on the pattern of regional Brandon EM, Wittmers LE Jr. 1988. Lead content
musculo-skeletal disease. Arthritis and and poisoning in Barbados slaves: historical,
Rheumatism 20:1019–1025. chemical, and biological evidence. In: Kiple KF,
Hadler NM, Gillings DB, Imbus HR, et al. 1978. editors. The African Exchange: Toward a
Hand structure and function in industrial Biological History of Black People. Durham, NC:
setting: influence of three patterns of Duke University Press; p. 140–166.
stereotyped, repetitive usage. Arthritis and Hanihara K. 1969. Mongoloid dental complex in
Rheumatism 21:210–220. the permanent dentition. Proceedings of the
Haensch S, Bianucci R, Signoli M, et al. 2010. VIIIth International Congress of
Distinct clones of Yersinia pestis caused the Anthropological and Ethnological Sciences,
Black Death. PloS Pathogens 6(10):1–8. 1968 I: 298–300. Tokyo, Japan: Science
Hakenbeck S, McManus E, Geisler H, Grupe G, Council of Japan.
O’Connell T. 2010. Diet and mobility in early Hanihara K. 1979. Dental traits in Ainu, Australian
Medieval Bavaria: a study of carbon and Aborigines, and New World populations. In:
nitrogen stable isotopes. American Journal of Laughlin WS, Harper AB, editors. The First
Physical Anthropology 143:235–249. Americans: Origins, Affinities, and
Halcrow S, Harris NJ, Beavan H, Buckley HR. Adaptations. New York, NY: Gustav Fischer;
2014. First bioarchaeological evidence of p. 125–134.
probable scurvy in Southeast Asia: Hanihara K. 1985. Origins and affinities of
multifactorial etiologies of vitamin C Japanese as viewed from cranial
deficiency in a tropical environment. measurements. In: Kirk R, Szathmáry E, editors.
International Journal of Paleopathology Out of Asia: Peopling of the Americas and the
5:63–71. Pacific. Canberra, Australia: The Journal of
Halcrow S, Tayles N. 2008. Stress near the start of Pacific History, Inc.; p. 104–112.
life? Localised enamel hypoplasia of the Hanihara K, Inoue N, Ito G, Kamegai T. 1981.
primary canine in late prehistoric mainland Microevolution and tooth to denture base
Southeast Asia. Journal of Archaeological discrepancy in Japanese dentition. Journal of
Science 35:2215–2222. the Anthropological Society of Nippon
Hales CN, Barker DJP, Clark PMS, et al. 1991. Fetal 89:63–70.
and infant growth and impaired glucose Hanihara T. 1994. Craniofacial continuity and
tolerance at age 64. British Medical Journal discontinuity of Far Easteners in the late
303:1019–1022. Pleistocene and Holocene. Journal of Human
Halffman CM, Scott GF, Pedersen PO. 1992. Evolution 27:417–441.
Palatine torus in the Greenlandic Norse. Hanihara T. 2013. Geographic structure of dental
American Journal of Physical Anthropology variation in the major human populations of
88:145–161. the world. In Scott GR, Irish JD, editors.
Hallberg L. 1981. Bioavailability of dietary iron in Anthropological Perspectives on Tooth
man. Annual Review of Nutrition 1:123–147. Morphology: Genetics, Evolution, Variation.
Hamperl H, Laughlin WS. 1959. Osteological Cambridge, UK: Cambridge University Press;
consequences of scalping. Human Biology p. 479–509.
31:80–89. Hanihara T, Ishida H. 2005. Metric dental
Hanawalt B. 1986. The Ties that Bound: Peasant variation of major human populations.
Families in Medieval England. Oxford, UK: American Journal of Physical Anthropology
Oxford University Press. 128:287–298.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 483

Hanihara T, Ishida H. 2009. Regional differences in Harding JE. 2001. The nutritional basis of the fetal
craniofacial diversity and the population origins of adult disease. International Journal
history of Jomon Japan. American Journal of of Epidemiology 30:15–23.
Physical Anthropology 139:311–322. Harding RM, Rösing FW, Sokal RR. 1989. Cranial
Hanihara T, Yoshida K, Ishida H. 2008. measurements do not support neolithisation of
Craniometric variation of the Ainu: an Europe by demic expansion. Homo 40:45–58.
assessment of differential gene flow from Hardinge MG, Swarner JB, Crooks H. 1965.
northeast Asia into northern Japan, Hokkaido. Carbohydrates in foods. Journal of American
American Journal of Physical Anthropology Dietetic Association 46:197–204.
137:283–293. Hardwick JL. 1960. The incidence and distribution
Hann JH. 1986. The use and processing of plants of caries throughout the ages in relation to the
by Indians of Spanish Florida. Southeastern Englishman’s diet. British Dental Journal
Archaeology 5:91–102. 108:9–17.
Hann JH. 1988. Apalachee: The Land Between the Hare PE, Fogel ML, Stafford TW Jr, Mitchell AD,
Rivers. Gainesville, FL: University Presses of Hoering TC. 1991. The isotopic composition of
Florida. carbon and nitrogen in individual amino acids
Hann JH. 1996. A History of the Timucua Indians isolated from modern and fossil proteins.
and Missions. Gainesville, FL: University Press Journal of Archaeological Science 18:277–292.
of Florida. Harman M, Molleson TI, Price JL. 1981. Burials,
Hansen JPH, Meldgaard J, Nordqvist J, editors. bodies and beheadings in Romano-British and
1991. The Greenland Mummies. Washington, Anglo-Saxon cemeteries. Bulletin of the British
DC: Smithsonian Institution Press. Museum of Natural History (Geology)
Hanson DB. 1988. Prehistoric mortuary practices 35:145–188.
and human biology. In: Butler BM, editor. Harper AB, Laughlin WS. 1982. Inquiries into the
Archaeological Investigations on the North peopling of the New World: development of
Coast of Rota, Mariana Islands. Micronesian ideas and recent advances. In: Spencer F,
Archaeological Survey, Report, 23, Southern editor. A History of American Physical
Illinois University at Carbondale, Center for Anthropology: 1930–1980. New York, NY:
Archaeological Investigations, Occasional Academic Press; p. 281–304.
Paper 8; p. 375–435. Harper KN, Zuckerman MK, Harper ML, Kingston
Hanson DB. 1990. Paleopathological observations JD, Armelagos GJ. 2011. The origin and
on human skeletal remains from Rota, Mariana antiquity of syphilis revisited: an appraisal of
Islands: epidemiological implications. Old World and pre-Columbian evidence for
Micronesica 2:349–362. treponemal infection. Yearbook of Physical
Hanson DB, Pietrusewsky M, editors. 1997. Special Anthropology 54:99–133.
issue: prehistoric skeletal biology in island Harper NK, Fox FC. 2008. Recent research in
ecosystems: current status of bioarchaeological Cypriot bioarchaeology. Bioarchaeology of the
research in the Marianas Archipelago. American Near East 2:1–38.
Journal of Physical Anthropology 104(3). Harris B. 1994. The height of schoolchildren in
Hara AT, Zero DT. 2010. The caries environment: Britain, 1900–1950. In: Komlos J, editors.
saliva, pellicle, diet, and hard tissue Stature, Living Standards, and Economic
ultrastructure. Dental Clinics of North America Development: Essays in Anthropometric
54:455–467. History. Chicago, IL: University of Chicago
Harbeck M, Seifert L, Hansch S, et al. 2013. Press; p. 25–38.
Yersinia pestis DNA from skeletal remains Harris EF. 1992. Laterality in human
from the 6th century AD reveals insights into odontometrics: analysis of a contemporary
Justinianic plague. PloS Pathogens 9(5): American White series. In: Lukacs JR, editor.
e10.1371/journal.ppat.1003349:1–8. Culture, Ecology and Dental Anthropology.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
484 References

Journal of Human Ecology, Special Issue 2: Harrod RP, Lienard P, Martin DL. 2012.
157–170. Deciphering violence: the potential of modern
Harris EF. 2008. Statistical application in dental ethnography to aid in the interpretation of
anthropology. In: Irish JD, Nelson GC, editors. archaeological populations. In: Martin DL,
Technique and Application in Dental Harrod RP, Pérez VR, editors. The
Anthropology. Cambridge, UK: Cambridge Bioarchaeology of Violence. Gainesville, FL:
University Press; p. 35–67. University Press of Florida; p. 63–80.
Harris EF, Bailit HL. 1980. The metaconule: a Hart DJ, Spector TD. 1993. The relationship of
morphologic and familial analysis of a molar obesity, fat distribution and osteoarthritis in
cusp in humans. American Journal of Physical women in the general population: the
Anthropology 53:349–358. Chingford Study. Journal of Rheumatology
Harris EF, Lease LR. 2005. Mesiodistal tooth crown 20:331–335.
dimension of the primary dentition: a Hartnady P, Rose JC. 1991. Abnormal tooth-loss
worldwide survey. American Journal of patterns among Archaic-period inhabitants of
Physical Anthropology 128:593–607. the lower Pecos region, Texas. In: Kelley MA,
Harris EF, Nweeia MT. 1980. Dental asymmetry as Larsen CS, editors. Advances in Dental
a measure of environmental stress in the Anthropology. New York, NY: Wiley-Liss;
Ticuna Indians of Columbia. American Journal p. 267–278.
of Physical Anthropology 53:133–142. Hartney PC. 1981. Tuberculosis lesions in a
Harris EF, Potter RH, Lin J. 2001. Secular trend in prehistoric population sample from southern
tooth size in urban Chinese assessed from two- Ontario. In: Buikstra JE, editor. Prehistoric
generation family data. American Journal of Tuberculosis in the Americas. Center for
Physical Anthropology 115:312–318. American Archeology, Scientific Papers, No. 5;
Harris EF, Rathbun TA. 1991. Ethnic differences in p. 141–160.
the apportionment of tooth sizes. In: Kelley Harvati K, Weaver TD. 2006. Human cranial
MA, Larsen CS, editors. Advances in Dental anatomy and the differential preservation of
Anthropology. New York, NY: Wiley-Liss; population history and climate signatures.
p. 121–142. Anatomical Record 288A:1225–1233.
Harris EF, Sjøvold T. 2004. Calculation of Smith’s Hassan FA. 1988. The Predynastic of Egypt.
Mean Measure of Divergence for intergroup Journal of World Prehistory 3:137–185.
comparisons using nonmetric data. Dental Hassett BR. 2014. Missing defects? A comparison
Anthropology 17:83–93. of microscopic and macroscopic approaches to
Harris HA. 1931. Lines of arrested growth in the identifying linear enamel hypoplasia.
long bones in childhood: the correlation of American Journal of Physical Anthropology
histological and radiographic appearances in 153:463–472.
clinical and experimental conditions. British Hastorf CA. 1990. The effect of the Inca state on
Journal of Radiology 4:561–588. Saysa agricultural production and crop
Harris HA. 1933. Bone Growth in Health and consumption. American Antiquity 50:262–290.
Disease: The Biological Principles Underlying Hastorf CA, Johannessen S. 1993. Pre-hispanic
the Clinical, Radiological and Histological political change and the role of maize in the
Diagnosis of Perversions of Growth and Disease central Andes of Peru. American
in the Skeleton. London, UK: Oxford Medical. Anthropologist 95:115–138.
Harrison RG, Katzenberg MA. 2003. Paleodiet Hatch JW, Geidel AA. 1985. Status-specific
studies using stable carbon isotopes from bone dietary variation in two world cultures. Journal
apatite and collagen: examples from southern of Human Evolution 14:469–476.
Ontario and San Nicolas Island, California. Hatch JW, Willey PS. 1974. Stature and status in
Journal of Anthropological Archaeology Dallas society. Tennessee Archaeologist
22:227–244. 30:107–131.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 485

Hatch JW, Willey PS, Hunt EE. 1983. Indicators of among Australian Aborigines. Atlantic
status-related stress in Dallas society: Highlands, NJ: Humanities Press.
transverse lines and cortical thickness in long Heaney RP. 1993. Nutritional factors in
bones. Midcontinental Journal of Archaeology osteoporosis. Annual Review of Nutrition
8:49–71. 13:287–316.
Hauser G, De Stefano GF. 1989. Epigenetic Heaney RP, Barger-Lux MJ, Davies KM, et al.
Variants of the Human Skull. Stuttgart, 1997. Bone dimensional change with age:
Germany: E. Schweizerbart’sche interactions of genetic, hormonal, and body
Verlagsbuchhandlung (Nägele u. Obermiller). size variables. Osteoporosis International
Hausman AJ. 1984. Holocene human evolution in 7:426–431.
southern Africa. In: Clark JD, Brandt SA, Hearne S. 1971. A Journey from Prince of Wales’s
editors. From Hunters to Farmers: The Causes Fort in Hudson’s Bay to the Northern Ocean.
and Consequences of Food Production in Edmonton, AB: M.G. Hurtic.
Africa. Berkeley, CA: University of California Heathcote GM. 1986. Exploratory Human
Press; p. 261–271. Craniometry of Recent Eskaleutian Regional
Havelkova P, Villotte S, Veleminsky P, Polacek L, Groups from the Western Arctic and Subarctic
Dobisikova M. 2011. Enthesopathies and of North America: A New Approach to
activity patterns in the early Medieval Great Population Historical Reconstruction. British
Moravian population: evidence of division of Archaeological Reports, International Series,
labor. International Journal of No. 301.
Osteoarchaeology 21:487–504. Heathcote GM. 1994. Population history
Haviland WA. 1967. Stature at Tikal, Guatemala: reconstruction, based on craniometry, I. The
implications for ancient Maya demography backtracking approach and initial results.
and social organization. American Antiquity Human Evolution 9:97–119.
32:316–325. Heaton THE, Vogel JC, Chevallarie G, Collett G. 1986.
Hawke SD, Davis JE. 1992. Seeds of Change: The Climatic influence on the isotopic composition of
Story of Cultural Exchange After 1492. Menlo bone nitrogen. Nature 322:822–823.
Park, CA: Addison-Wesley. Hedges REM. 2002. Bone diagenesis: an overview
Hawkes K, O’Connell JF, Blurton Jones NG. 2001. of processes. Archaeometry 44:319–328.
Hadza meat sharing. Evolution and Human Hedges REM, Clement JG, Thomas CD, O’Connell
Behavior 22:113–142. TC. 2007. Collagen turnover in the adult
Hawkey DE. 1998. Disability, compassion and the femoral mid-shaft: modeled from
skeletal record: using musculoskeletal stress anthropogenic radiocarbon tracer
markers (MSM) to construct an osteobiography measurements. American Journal of Physical
from early New Mexico. International Journal Anthropology 133:808–816.
of Osteoarchaeology 8:326–340. Hedges REM, Reynard LM. 2007. Nitrogen isotopes
Hawkey DE, Merbs CF. 1995. Activity-induced and the trophic level of humans in
musculoskeletal stress markers (MSM) and archaeology. Joural of Archaeological Sciences
subsistence strategy changes among ancient 34:1240–1251.
Hudson Bay Eskimos. International Journal of Hedman K, Hargrave E, Ambrose SH. 2002. Inter-
Osteoarchaeology 5:324–338. and intra-site comparisons of Mississippian
Hawkey DE, Street SR. 1992. Activity-induced diet in the American Bottom: results of recent
stress markers in prehistoric human remains stable isotope analyses of bone collagen and
from the eastern Aleutian Islands. American apatite. Midcontinental Journal of Archaeology
Journal of Physical Anthropology, Supplement 27:237–271.
14:89. Heider KG. 1979. Grand Valley Dani: Peaceful
Hayden B. 1979. Paleolithic Reflections: Lithic Warriors. New York, NY: Holt, Rinehart and
Technology and Ethnographic Excavation Winston.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
486 References

Heithersay G. 1959. A dental survey of the South Asian Archaeology 1995. New Delhi,
aborigines at Haast’s Bluff, Central Australia. India: Oxford-IBH.
Medical Journal of Australia 1:721–729. Hemphill BE, Christensen AF, Mustafakulov SI.
Helmuth J, Pendergast DM. 1986–1987. Lamanai 1998. Biological adaptations and affinities of
Tomb N9–58/1: analysis of the skeletal Bronze Age Bactrians: II. Dental morphology.
evidence. Ossa 13:109–118. In: Lukacs JR, editor. Human Dental
Hemphill BE. 1991. Tooth Size Apportionment Development, Morphology, and Pathology.
among Contemporary Indians: An Analysis of University of Oregon Anthropological Papers,
Caste, Language, and Geography. PhD No. 54; p. 51–77.
Dissertation, University of Oregon, Eugene, OR. Hemphill BE, Larsen CS, editors. 2010.
Hemphill BE. 1992. An Osteological Analysis of Understanding Prehistoric Lifeways in the
the Human Remains from Malheur Lake, Great Basin Wetlands: Bioarchaeological
Oregon. (3 volumes). U.S. Department of the Reconstruction and Interpretation. Salt Lake
Interior, Fish and Wildlife Service, Cultural City, UT: University of Utah Press. (Paperback
Resource Series, No. 6. edition of Hemphill & Larsen, 1999.)
Hemphill BE. 1998. Biological affinities and Hemphill BE, Lukacs JR, Kennedy KAR. 1991.
adaptations of Bronze Age Bactrians: III. An Biological adaptations and affinities of
initial craniometric assessment. American Bronze Age Harappans. In: Meadow R, editor.
Journal of Physical Anthropology Harappa Excavations 1986–1990: A
106:329–348. Multidisciplinary Approach to Third
Hemphill BE. 1999a. Biological affinities and Millennium Urbanism. Madison, WI: Prehistory
adaptations of Bronze Age Bactrians: IV. Press; p. 137–182.
A craniometric investigation of Bactrian Hemphill BE, Mallory JP. 2004. Horse-mounted
origins. American Journal of Physical invaders from the Russo-Kazakh Steppe or
Anthropology 108:173–192. agricultural colonists from western Central
Hemphill BE. 1999b. Foreign elites from the Oxus Asia? A craniometric investigation of the
civilization? A craniometric study of Bronze Age settlement of Xinjiang. American
anomalous burials from Bronze Age Tepe Journal of Physical Anthropology
Hissar. American Journal of Physical 124:199–222.
Anthropology 110:421–434. Henderson CY, Cardoso FA, editors. 2013. Special
Hemphill BE. 2008. Dental pathology prevalence Issue, Entheseal Changes and Occupation:
and pervasiveness at Tepe Hissar: statistical Technical and Theoretical Advances and Their
utility for investigating inter-relationships Applications. International Journal of
between wealth, gender and status. In Irish JD, Osteoarchaeology 23:127–251.
Nelson GC, editors. Technique and Application Henke W. 1984. On cranial morphology and
in Dental Anthropology. Cambridge, UK: climate. Current Anthropology 25:533–534.
Cambridge University Press; p. 178–215. Henneberg M. 1984. Comment. Current
Hemphill BE. 2010. Wear and tear: osteoarthritis Anthropology 25:321–322.
as an indicator of mobility among Great Basin Henneberg M, Henneberg R, Carter J. 1992. Health
hunter-gatherers. In: Hemphill BE, Larsen CS, in colonial Metaponto. National Geographic
editors. Understanding Prehistoric Lifeways in Research & Exploration 8:446–459.
the Great Basin Wetlands: Bioarchaeological Henneberg M, Henneberg RJ. 1994.
Reconstruction and Interpretation. Salt Lake Treponematosis in an ancient Greek colony of
City, UT: University of Utah Press; p. 241–289. Metaponto, southern Italy, 580–250 BCE. In:
Hemphill BE, Christensen AF, Mustafakulov SI. Dutour O, Pálfi G, Bérato JP, editors. L’Origine
1996. East meets West: a diachronic analysis of de la Syphilis en Europe: Avant ou Après 1493?
Bronze Age biological interactions across the Toulon, France: Centre Archeologique du Var,
Indo-Iranian borderlands. In: Allchin B, editor. Éditions Errance; p. 92–98.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 487

Henneberg RJ. 1996. Dental health of an urban Hiebert FT, Lamberg-Karlovsky CC. 1992. Central
population in 7th–2nd century B.C. Metaponto, Asia and the Indo-Iranian Borderlands. Iran
south Italy. American Journal of Physical 30:1–15.
Anthropology Supplement 22:122. Higgins RL, Haines MR, Walsh L, Sirianni JE.
Henry CJK, Ulijaszek SJ, editors. 1996. Long-Term 2002. The poor in the mid-nineteenth-century
Consequences of Early Environment: Growth, northeastern United States: evidence from the
Development and the Lifespan Developmental Monroe County Almshouse, Rochester, York.
Perspective. Cambridge, UK: Cambridge In: Steckel RH, Rose JC, editors. The Backbone
University Press. of History: Health and Nutrition in the Western
Herring DA, Saunders SR, Katzenberg MA. 1998. Hemisphere. New York, NY: Cambridge
Investigating the weaning process in past University Press; p. 162–184.
populations. American Journal of Physical Hildebolt CF, Elvin-Lewis M, Molnar S, et al. 1989.
Anthropology 105:425–439. Caries prevalence among geochemical regions
Herring SW. 1993. Epigenetic and functional of Missouri. American Journal of Physical
influences on skull growth. In: Hanken J, Hall Anthropology 78:79–92.
BK, editors. The Skull. Chicago, IL: University Hildebolt CF, Molnar S. 1991. Measurement and
of Chicago Press; p. 153–206. description of periodontal disease in
Herrscher W, Le Bras-Goude G. 2010. Southern anthropological studies. In: Kelley MA, Larsen
French Neolithic populations: isotopic CS, editors. Advances in Dental Anthropology.
evidence for regional specificities in New York, NY: Wiley-Liss; p. 225–240.
environment and diet. American Journal of Hildebolt CF, Molnar S, Elvin-Lewis M, McKee JK.
Physical Anthropology 141:259–272. 1988. The effect of geochemical factors on
Hershkovitz I, Donoghue HD, Minnikin DE, prevalences of dental diseases for prehistoric
et al. 2008. Detection and molecular inhabitants of the State of Missouri.
characterization of 9000-year-old American Journal of Physical Anthropology
Mycobacterium tuberculosis from a Neolithic 75:1–14.
settlement in the eastern Mediterranean. Hill K, Hurtado AM. 1995. Ache Life History: The
PLoS ONE 3:1–6. Ecology and Demography of a Foraging People.
Hershkovitz I, Galili E. 1990. 8000-years old New York, NY: Aldine de Gruyter.
human remains on the sea floor near Atlit, Hill KR, Walker RS, Bozicevic M, et al. 2011.
Israel. Human Evolution 5:319–358. Co-residence patterns in hunter-gatherer
Hershkovitz I, Gopher A. 2008. Demographic, societies show unique human social structure.
biological and cultural aspects of the Neolithic Science 331:1286–1289.
revolution: a view from the southern Levant. Hill MC. 1981. Analysis, synthesis, and
In: Bocquet-Appel J-P, Bar-Yosef O, editors. interpretation of the skeletal material
The Neolithic Demographic Transition and Its excavated for the Gainesville section of the
Consequences. New York, NY: Springer; Tennessee-Tombigbee Waterway. In: Caddell
p. 441–479. GM, Woodrick A, Hill MC, editors. Biocultural
Hewlett BS. 1991. Demography and childcare in Studies in the Gainesville Lake Area. University
preindustrial societies. Journal of of Alabama, Office of Archaeological Research,
Anthropological Research 47:1–37. Report of Investigations, No.14; p. 209–334.
Hewlett BS, van de Koppel JMH, van de Koppel M. Hillier RJ, Craig GT. 1992. Human dental enamel
1982. Exploration ranges of Aka pygmies in the determination of health patterns in
of the Central African Republic. Man children. In: Goodman AH, Capasso LL, editors.
17:418–430. Recent Contributions to the Study of Enamel
Heyberger L. 2007. Toward an anthropometric Developmental Defects. Journal of
history of provincial France, 1780–1920. Paleopathology, Monographic Publications,
Economics and Human Biology 5:229–254. No. 2; p. 381–390.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
488 References

Hillson S. 1979. Diet and dental disease. World Hinton RJ. 1981c. Changes in articular eminence
Archaeology 11:147–162. morphology with dental function. American
Hillson S. 1996. Dental Anthropology. Cambridge, Journal of Physical Anthropology 54:439–455.
UK: Cambridge University Press. Hinton RJ. 1981d. Form and patterning of anterior
Hillson S. 2014. Tooth Development in Human tooth wear among aboriginal human groups.
Evolution and Bioarchaeology. Cambridge, UK: American Journal of Physical Anthropology
Cambridge University Press. 54:555–564.
Hillson SW. 2005. Teeth, Second Edition. Hinton RJ. 1982. Differences in interproximal and
Cambridge, UK: Cambridge University Press. occlusal tooth wear among prehistoric
Hillson SW. 2008. The current state of dental Tennessee Indians: implications for
decay. In: Irish JD, Nelson GC, editors. masticatory function. American Journal of
Technique and Application in Dental Physical Anthropology 57:103–115.
Anthropology. Cambridge, UK: Cambridge Hinton RJ. 1983. Relationships between
University Press; p. 111–135. mandibular joint size and craniofacial size in
Hillson SW, Bond S. 1997. Relationship of enamel human groups. Archives of Oral Biology
hypoplasia to the pattern of tooth crown 28:37–43.
growth: a discussion. American Journal of Hinton RJ. 1990. Myotomy of the lateral
Physical Anthropology 104:89–104. pterygoid muscle and condylar cartilage
Hillson S, Grigson C, Bond S. 1998. Dental growth. European Journal of Orthodontics
defects of congenital syphilis. American 12:370–379.
Journal of Physical Anthropology Hinton RJ, Carlson DS. 1979. Temporal changes in
107:25–40. human temporomandibular joint size and
Hillson SW, Larsen CS, Boz B, et al. 2013. The shape. American Journal of Physical
human remains. I. Interpreting community Anthropology 50:325–334.
structure, health, and diet in Neolithic Hinton RJ, Smith MO, Smith FH. 1980. Tooth size
Çatahöyük. In: Hodder I, editor. Humans and changes in prehistoric Tennessee Indians.
Landscapes of Çatalhöyük. Los Angeles, CA: Human Biology 52:229–245.
Cotsen Institute of Archaeology Press; Hirata K. 1990. Secular trend and age distribution
p. 339–396. of cribra orbitalia in Japanese. Human
Himes JH. 1978. Bone growth and development in Evolution 5:375–385.
protein-calorie malnutrition. World Review of Hirata K, Oku C, Morimoto I. 2000. A case of
Nutrition and Dietetics 28:143–187. leprosy in a Medieval Japanese.
Himes JH. 1979. Secular Changes in Body Anthropological Science 108:144.
Proportions and Composition. Monographs of Hobson KA, Alisauskas RT, Clark RG. 1993.
Society for Child Development, No. 44. Stable-nitrogen isotope enrichment in avian
Himes JH, Martorell R, Habicht J-P, Malina RM, tissues due to fasting and nutritional stress:
Klein RE. 1975. Patterns of cortical bone implications for isotope analysis of diet.
growth in moderately malnourished preschool Condor 95:388–394.
children. Human Biology 47:337–350. Hodges DC. 1987. Health and agricultural
Hinton RJ. 1981a. Temporomandibular joint size intensification in the prehistoric valley of
alterations in prehistoric Tennessee Indians. Oaxaca, Mexico. American Journal of Physical
Tennessee Anthropologist 6:89–111. Anthropology 73:323–332.
Hinton RJ. 1981b. Form and function in the Hodges DC. 1989. Agricultural Intensification and
temporomandibular joint. In: Carlson DS, Prehistoric Health in the Valley of Oaxaca,
editor. Craniofacial Biology. University of Mexico. University of Michigan, Memoirs of
Michigan Center for Human Growth and the Museum of Anthropology, No. 22.
Development, Craniofacial Growth Series, Hodder I. 2006. Çatalhöyük: The Leopard’s Tale.
Monograph, No. 10; p. 37–60. London, UK: Thames and Hudson.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 489

Hodder I, editor. 2013. Humans and Landscapes of Holt BM, Formicola V. 2008. Hunters of the
Çatalhöyük. Los Angeles, CA: Cotsen Institute Ice Age: the biology of Upper Paleolithic
of Archaeology. people. Yearbook of Physical Anthropology
Hohl TH. 1983. Masticatory muscle transposition 51:70–90.
in primates: effects on craniofacial growth. Holt B, Ruff C, Niskanen M, et al. 2012.
Journal of Maxillo-Facial Surgery 11:149–156. Postcranial robusticity trends in Europe across
Hojo T. 1989. Dietary differences and microwear the last 30,000 years. American Journal of
on the teeth of late Stone Age and early Physical Anthropology Supplement 54:167.
modern people from western Japan. Scanning Holtzman J. 2009. Uncertain Tasks: Memory,
Microscopy 3:623–628. Ambivalence, and the Politics of Eating in
Holick MF. 1998. Perspective on the impact of Samburu, Northern Kenya. Berkeley, CA:
weightlessness on calcium and bone University of California Press.
metabolism. Bone 22(S5):105S–111S. Homan BT. 1977. Changing periodontal status in a
Hollimon SE. 1991. Health consequences of changing environment. Journal of Dental
divisions of labor among the Chumash Indians Research 56(Special Issue C):C46–C54.
of southern California. In: Walde D, Willows Homes Hogue S, Melsheimer R. 2008. Integrating
ND, editors. The Archaeology of Gender. dental microwear and isotopic analyses to
Archaeological Association of the University of understand dietary change in East-Central
Calgary, Proceedings of the 22nd Annual Mississippi. Journal of Archaeological Science
Chacmool Conference; p. 462–469. 35:228–238.
Hollimon SE, Owsley DW. 1994. Osteology of the Hoogland MLP, Hofman CL, Panhuysen GAM.
Fay Tolton site: implications for warfare during 2010. Interisland dynamics: evidence for
the Initial Middle Missouri variant. In: Owsley human mobility at the site of Anse a la Gourde,
DW, Jantz RL, editors. Skeletal Biology in the Guadeloupe. In: Fitzpatrick SM, Ross AH,
Great Plains: Migration, Warfare, Health, and editors. Island Shores, Distant Pasts:
Subsistence. Washington, DC: Smithsonian Archaeological and Biological Approaches to
Institution Press; p. 345–353. the Pre-Columbian Settlement of the Caribbean.
Hollister MC, Weintraub JA. 1993. The association Gainesville, FL: University Press of Florida;
of oral status with systemic health, quality of p. 148–162.
life, and economic productivity. Journal of Hooton EA. 1918. On certain Eskimoid characters
Dental Education 57:901–912. in Icelandic skulls. American Journal of
Holloway PJ, Shaw JH, Sweeney EA. 1961. Effects Physical Anthropology 1:53–76.
of various sucrose:casein ratios in purified diets Hooton EA. 1930. The Indians of Pecos Pueblo:
on the teeth and supporting structures of rats. A Study of Their Skeletal Remains. New Haven,
Archives of Oral Biology 3:185–200. CT: Yale University Press.
Holman DJ, Yamaguchi K. 2005. Longitudinal Hooton EA. 1933. Racial types in America and
analysis of deciduous tooth emergence: IV. their relation to Old World types. In: Jenness D,
Covariate effects in Japanese children. editor. The American Aborigines. New York,
American Journal of Physical Anthropology NY: Russell and Russell; p. 131–163.
126:352–358. Hooton EA. 1940. Skeletons from the Cenote of
Holmes MA, Ruff CB. 2011. Dietary effects on Sacrifice at Chichen Itza. In: Hay CL, Linton RL,
development of the human mandibular corpus. Lothrop SK, Shapiro HL, Vaillant CG, editors.
American Journal of Physical Anthropology The Maya and Their Neighbors. New York, NY:
145:615–628. Dover; p. 270–280.
Holt BM. 2003. Mobility in Upper Paleolithic and Hooton EA, Dupertuis CW. 1951. Age Changes and
Mesolithic Europe: evidence from the lower Selective Survival in Irish Males. Studies in
limb. American Journal of Physical Physical Anthropology, 2. New York, NY:
Anthropology 122:200–215. American Association of Physical

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
490 References

Anthropologists and Wenner-Gren Foundation American Journal of Physical Anthropology


for Anthropological Research. 59:263–269.
Hopewell PC. 1994. Overview of clinical Howell TL, Kintigh KW. 1996. Archaeological
tuberculosis. In: Bloom BR, editor. identification of kin groups using mortuary
Tuberculosis: Pathogenesis, Protection, and and biological data: an example from the
Control. Washington, DC: American Society for American Southwest. American Antiquity
Microbiology Press; p. 25–46. 61:537–554.
Hoppa RD. 1992. Evaluating human skeletal Howells WW. 1960. Estimating population
growth: an Anglo-Saxon example. numbers through archaeological and skeletal
International Journal of Osteoarchaeology remains. In: Heizer F, Cook SF, editors. The
2:275–288. Application of Quantitative Methods in
Hoppa RD. 2002. Paleodemography: looking back Archaeology. New York, NY: Viking Fund;
and thinking ahead. In: Hoppa RD, Vaupel JW, p. 158–185.
editors. Paleodemography: Age Distributions Howells WW. 1966. The Jomon Population of
from Skeletal Samples. Cambridge, UK: Japan. Harvard University, Papers of the
Cambridge University Press; p. 9–28. Peabody Museum of Archaeology and
Hoppa RD, FitzGerald CM, editors. 1999. Human Ethnology, 57.
Growth in the Past: Studies from Bones and Howells WW. 1973. Cranial Variation in Man:
Teeth. Cambridge, UK: Cambridge University A Study by Multivariate Analysis of Patterns of
Press. Difference among Recent Human Populations.
Hoppa RD, Vaupel JW, editors. 2002. Harvard University, Papers of the Peabody
Paleodemography: Age Distributions from Museum of Archaeology and Ethnology, 67.
Skeletal Samples. Cambridge, UK: Cambridge Howells WW. 1986. Physical anthropology of the
University Press. prehistoric Japanese. In: Pearson RJ, Barnes Gl,
Horowitz S, Shapiro HH. 1951. Modifications of Hutterer KL, editors. Windows on the Japanese
mandibular architecture following removal of Past. Ann Arbor, MI: Center for Japanese
the temporalis in the rat. Journal of Dental Studies, University of Michigan; p. 85–90.
Research 30:276–280. Howells WW. 1989. Skull Shapes and the Map:
Horowitz S, Shapiro HH. 1955. Modification of Craniometric Analyses in the Dispersion of
the skull and jaw architecture following Modern Homo. Harvard University, Papers of
removal of the masseter muscle in the rat. the Peabody Museum of Archaeology and
American Journal of Physical Anthropology Ethnology, 79.
13:301–308. Howells WW, Bleibtreu HK. 1970. Hutterite Age
Hoshina H. 1980. Spondylolysis in athletes. The Differences in Body Measurements. Harvard
Physician and Sportsmedicine 8:75–79. University, Papers of the Peabody Museum of
Hough AJ Jr. 2001. Pathology of osteoarthritis. In: American Archaeology and Ethnology, 57(2).
Koopman WJ, editor. Arthritis and Allied Howitt R. 1998. Recognition, respect and
Conditions, Fourteenth Edition. Philadelphia, reconciliation: steps towards decolonization?
PA: Lippincott, Williams and Wilkins; p. 69–99. Australian Aboriginal Studies, p. 28–34.
Hough AJ, Sokoloff L. 1989. Pathology of Howland MR, Corr LT, Young SM, et al. 2003.
osteoarthritis. In: McCarty DJ, editor. Arthritis Expression of the dietary isotope signal in the
and Allied Conditions, Eleventh Edition. compound-specific δ13C values of pig bone
Philadelphia, PA: Lea & Febiger; p. 1571–1594. lipids and amino acids. International Journal of
Houston CS, Zaleski WA. 1967. The shape of Osteoarchaeology 13:54–65.
vertebral bodies and femoral necks in relation Hoyme LE, Bass WM. 1962. Human skeletal
to activity. Radiology 89:59–66. remains from the Tollifero (Ha6) and
Howell N. 1982. Village composition implied by a Clarksville (Mc14) sites, John H. Kerr Reservoir
paleodemographic life table: the Libben site. Basin, Virginia. In: Miller CF, editor.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 491

Archeology of the John H. Kerr Reservoir Basin, Hrdlička A. 1941. Diseases of and Artifacts on
Roanoke River, Virginia-North Carolina. Skull and Bones from Kodiak Island.
Bureau of American Ethnology, Bulletin, Smithsonian Miscellaneous Collections, 101(4).
No. 182; p. 329–400. Hrdlička A. 1944. The Anthropology of Kodiak
Hrdlička A. 1909. Tuberculosis among Certain Island. Philadelphia, PA: Wistar Institute of
Indian Tribes of the United States. Bureau of Anatomy and Biology.
American Ethnology, Bulletin, No. 42. Hrdlička A. 1945. The Aleutian and Commander
Hrdlička A. 1910a. An Ancient Sepulchre at San Islands and Their Inhabitants. Philadelphia,
Juan Teotihuacan, with Anthropological Notes PA: Wistar Institute of Anatomy and Biology.
on the Teotihuacan People. XVII Congreso Hsiang SM, Burke M, Miguel E. 2013. Quantifying
Internacional de Americanistas (Buenos Aires), the influence of climate on human conflict.
Appendix 3–7. Science 341:1213–1225.
Hrdlička A. 1910b. Contribution to the Hubbard A, Guatelli-Steinberg D, Sciulli PW.
Anthropology of Central and Smith Sound 2009. Under restrictive conditions, can the
Eskimos. Anthropological Papers of the widths of linear enamel hypoplasias be used as
American Museum of Natural History, relative indicators of stress episode duration?
No. 5 (part 2). American Journal of Physical Anthropology
Hrdlička A. 1912. Remains in Eastern Asia of the 138:177–189.
Race that Peopled America. Smithsonian Hubbe M, Hanihara T, Harvati K. 2009. Climate
Miscellaneous Collections, 60(16). signatures in the morphological differentiation
Hrdlička A. 1914. Anthropological Work in Peru in of worldwide modern human populations.
1913, with Notes on the Pathology of the Anatomical Record 292:1720–1733.
Ancient Peruvians. Smithsonian Miscellaneous Hubbe M, Harvati K, Neves W. 2011.
Collections, 61(18). Paleoamerican morphology in the context of
Hrdlička A. 1916. Physical Anthropology of the European and East Asian Late Pleistocene
Lenape or Delawares and the Eastern Indians in variation: implications for human dispersion
General. Bureau of American Ethnology into the New World. American Journal of
Bulletin, No. 62. Physical Anthropology 144:442–453.
Hrdlička A. 1922a. The Anthropology of Florida. Hubbe M, Neves WA, Harvati K. 2010. Testing
Publications of the Florida State Historical evolutionary and dispersion scenarios for
Society, 1. the settlement of the New World. PLoS ONE
Hrdlička A. 1922b. The causes of malocclusion. 5:e11105.
Dental Cosmos 64:489–497. Hubbe M, Torres-Rouff C, Neves WA, et al. 2012.
Hrdlička A. 1932. The principal dimensions, Dental health in northern Chile’s Atacama
absolute and relative, of the humerus in the Oases: evaluating the Middle Horizon
White race. American Journal of Physical (AD 500–1000) impact on local diet.
Anthropology 6:431–450. American Journal of Physical Anthropology
Hrdlička A. 1935. Ear Exostoses. Smithsonian 148:62–72.
Miscellaneous Collections, 93(6). Hudson C. 1976. The Southeastern Indians.
Hrdlička A. 1936. Growth during adult life. Knoxville, TN: University of Tennessee Press.
Proceedings of the American Philosophical Hudson EH. 1958. Non-Venereal Syphilis:
Society 76:847–897. A Sociological and Medical Study of Bejel.
Hrdlička A. 1940a. Ritual ablation of front teeth in Edinburgh, UK: E. & S. Livingstone.
Siberia and America. Smithsonian Hudson EH. 1965. Treponematosis and man’s
Miscellaneous Collections, 99(1). social evolution. American Anthropologist
Hrdlička A. 1940b. Mandibular and maxillary 67:885–901.
hyperostoses. American Journal of Physical Hughes T, Dempsey P, Richards L, Townsend G.
Anthropology 27:1–67. 2000. Genetic analysis of deciduous tooth size

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
492 References

in Australian twins. Archives of Oral Biology enamel. Proceedings of the National Academy
45:997–1004. of Sciences 105:6834–6839.
Hughes TE, Townsend GC. 2013. Twin and family Humphreys HC, King H, editors. 1981. Mortality
studies of human dental crown morphology: and Immortality: The Anthropology and
genetic, epigenetic, and environmental Archaeology of Death. London, UK: Academic
determinants of the modern human dentition. Press.
In: Scott GR, Irish JD, editors. Anthropological Hunt EE Jr. 1960. The continuing evolution of
Perspectives on Tooth Morphology: Genetics, modern man. Cold Spring Harbor Symposia
Evolution, Variation. Cambridge, UK: on Quantitative Biology 24:245–254.
Cambridge University Press; p. 31–68. Hunt EE Jr. 1961. Malocclusion and
Huiskes R. 1982. On the modelling of long bones civilization. American Journal of Orthodontics
in structural analysis. Journal of Biomechanics 47:406–422.
15:65–69. Hunt RJ, Drake CW, Beck JD. 1992. Streptococcus
Hulse FS. 1941. The people who lived at Irene: mutans, lactobacilli, and caries experience in
physical anthropology. In: Caldwell J, McCann older adults. Special Care in Dentistry
C. Irene Mound Site, Chatham County, Georgia. 12:149–152.
Athens, GA: University of Georgia Press; Hunter J, Guatelli-Steinberg D, Weston TC, Durner
p. 57–68. R, Betsinger T. 2010. Model of tooth
Hummert JR. 1983. Cortical bone growth and morphogenesis predicts Carabelli cusp
dietary stress among subadults from Nubia’s expression, size, and symmetry in humans.
Batn el Hajar. American Journal of Physical PLoS ONE 5:e11844.
Anthropology 62:167–176. Hurlbut SA. 2000. The taphonomy of cannibalism:
Hummert JR, Van Gerven DP. 1982. Tetracycline- a review of anthropogenic bone modification
labeled human bone from a Medieval in the American Southwest. International
population in Nubia’s Batn el Hajar Journal of Osteoarchaeology 10:4–26.
(550–1450 A.D.). Human Biology 54:355–371. Hussain K, Wijetunge DB, Grubnic S, Jackson IT.
Hummert JR, Van Gerven DP. 1983. Skeletal 1994. A comprehensive analysis of craniofacial
growth in a Medieval population from trauma. Journal of Trauma 36:34–47.
Sudanese Nubia. American Journal of Physical Huss-Ashmore R. 1981. Bone growth and
Anthropology 60:471–478. remodeling as a measure of nutritional stress.
Humphrey LT. 2003. Linear growth variation in In: Martin DL, Bumsted MP, editors.
the archaeological record. In: Thompson JL, Biocultural Adaptation: Comprehensive
Krovitz GE, Nelson AJ, editors. Patterns of Approaches to Skeletal Analysis. University of
Growth and Development in the Genus Homo. Massachusetts, Department of Anthropology,
Cambridge, UK: Cambridge University Press; Research Reports, No. 20; p. 84–95.
p. 144–169. Huss-Ashmore R, Goodman AH, Armelagos GJ.
Humphrey LT, Bocaege E. 2008. Tooth evulsion in 1982. Nutritional inference from
the Maghreb: chronological and geographical paleopathology. In: Schiffer MB, editor.
patterns. African Archaeological Review Advances in Archaeological Method and
25:109–123. Theory. New York, NY: Academic Press;
Humphrey LT, De Groote I, Morales J, et al. 2014. p. 395–474.
Earliest evidence for caries and exploitation Hutchinson DL. 1991. Postcontact Native
of starchy plant foods in Pleistocene American Health and Adaptation: Assessing
hunter-gatherers from Morocco. Proceedings the Impact of Introduced Diseases in Sixteenth-
of the National Academy of Sciences Century Gulf Coast Florida. PhD Dissertation,
111:954–959. University of Illinois, Urbana, IL.
Humphrey LT, Dean MC, Jeffries TE, Penn M. Hutchinson DL. 1993. Treponematosis in regional
2008. Unlocking evidence of early diet from and chronological perspective from central

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 493

Gulf Coast Florida. American Journal of Hutchinson DL, Larsen CS, Choi I. 1994. Stressed
Physical Anthropology 92:249–261. to the max: physiological perturbation in the
Hutchinson DL. 1996. Brief encounters: Tatham Krapina Neandertals. American Journal of
Mound and the evidence for Spanish and Physical Anthropology S18:110.
Native American confrontation. International Hutchinson DL, Larsen CS, Choi I. 1997. Stressed
Journal of Osteoarchaeology 6:51–65. to the max? Physiological perturbation in the
Hutchinson DL. 2002. Foraging, Farming, and Krapina Neandertals. Current Anthropology
Coastal Biocultural Adaptation in Late 38:904–914.
Prehistoric North Carolina. Gainesville, FL: Hutchinson DL, Larsen CS, Norr L. 2000.
University Press of Florida. Agricultural melodies and alternative
Hutchinson DL. 2004. Bioarchaeology of the harmonies in Florida and Georgia. In:
Florida Gulf Coast: Adaptation, Conflict, and Lambert PM, editor. Bioarchaeological
Change. Gainesville, FL: University Press of Studies of Life in the Age of Agriculture.
Florida. Tuscaloosa, AL: University of Alabama Press;
Hutchinson DL. 2006. Tatham Mound and the p. 116–133.
Bioarchaeology of European Contact: Disease Hutchinson DL, Larsen CS, Schoeninger MJ, Norr
and Depopulation in Central Gulf Coast L. 1998. Regional variation in the pattern of
Florida. Gainesville, FL: University Press of maize adoption and use in Florida and Georgia.
Florida. American Antiquity 63:397–416.
Hutchinson DL, Humphrey J. 2002. Macroscopic Hutchinson DL, Larsen CS, Williamson MA, Green
characteristics of hacking trauma. Journal of Clow VD, Powell ML. 2005. Temporal and
Forensic Sciences 46:228–233. spatial variation in the patterns of
Hutchinson DL, Larsen CS. 1988. Determination treponematosis in Georgia and Florida. In:
of stress episode duration from linear Powell ML, Cook DC, editors. The Myth of
enamel hypoplasias: a case study from Syphilis: The Natural History of
St. Catherines Island, Georgia. Human Biology Treponematosis in North America. Gainesville,
60:93–110. FL: University Press of Florida; p. 92–116.
Hutchinson DL, Larsen CS. 1990. Stress and Hutchinson DL, Mitchem JM. 1996. The Weeki
lifeway change: the evidence from enamel Wachee Mound, an early contact period
hypoplasias. In: Larsen CS, editor. The mortuary locality in Hernando County, west-
Archaeology of Mission Santa Catalina de central Florida. Southeastern Archaeology
Guale: 2. Biocultural Interpretations of a 15:47–65.
Population in Transition. Anthropological Hutchinson DL, Norr L. 1994. Late prehistoric and
Papers of the American Museum of Natural early historic diet in Gulf Coast Florida. In:
History, No. 68; p. 50–65. Larsen CS, Milner GR, editors. In the Wake of
Hutchinson DL, Larsen CS. 1995. Physiological Contact: Biological Responses to Conquest.
stress in the prehistoric Stillwater Marsh: New York, NY: Wiley-Liss; p. 9–20.
evidence of enamel defects. In: Larsen CS, Kelly Hutchinson DL, Norr L. 2006. Nutrition and health
RL, editors. Bioarchaeology of the Stillwater at contact in late prehistoric central Gulf Coast
Marsh: Prehistoric Human Adaptation in the Florida. American Journal of Physical
Western Great Basin. Anthropological Papers Anthropology 129:375–386.
of the American Museum of Natural History, Hutchinson DL, Norr L, Teaford MF. 2007. Outer
No. 77; p. 81–95. coast forager and inner coast farmers in late
Hutchinson DL, Larsen CS. 2001. Enamel prehistoric North Carolina. In: Cohen MN,
hypoplasia and stress in La Florida. In: Larsen Crane-Kramer GMM, editors. Ancient Health:
CS, editor. Bioarchaeology of Spanish Florida: Skeletal Indicators of Agricultural and
The Impact of Colonialism. Gainesville, FL: Economic Intensification. Gainesville, FL:
University Press of Florida; p. 181–206. University Press of Florida; p. 52–64.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
494 References

Hutchinson DL, Richman R. 2006. Regional, social, editors. Prehistoric Hunter-Gatherers in Japan:
and evolutionary perspectives on treponemal New Research Methods. University of Tokyo
infection in the southeastern United States. Museum Bulletin, No. 27; p. 163–198.
American Journal of Physical Anthropology Irei K, Doi N, Fukumine T, et al. 2008. Dental
129:544–558. disease of human skeletal remains from the
Hutchinson DL, Weaver DS. 1998. Two cases of early modern period of Kumejima Island,
facial involvement in probable treponemal Okinawa, Japan. Anthropological Science
infection from late prehistoric coastal North 116:149–159.
Carolina. International Journal of Irish JD. 2000. The Iberomaurusian enigma: North
Osteoarchaeology 8:444–453. African progenitor or dead end? Journal of
Hylander WL. 1977. The adaptive significance of Human Evolution 39:393–410.
Eskimo craniofacial morphology. In: Dahlberg Irish JD. 2005. Population continuity vs.
AA, Graber TM, editors. Orofacial Growth and discontinuity revisited: dental affinities among
Development. The Hague, the Netherlands: late Paleolithic through Christian-era Nubians.
Mouton; p. 129–169. American Journal of Physical Anthropology
Hylander WL, Picq PG, Johnson KR. 1991. 128:520–535.
Function of the supraorbital region of primates. Irish JD. 2006. Who were the ancient Egyptians?
Archives of Oral Biology 36:273–281. Dental affinities among Neolithic through
Ibanez-Gimeno P, De Esteban-Trivigno S, Jordana Postdynastic peoples. American Journal of
X, et al. 2013. Functional plasticity of the Physical Anthropology 129:529–543.
human humerus: shape, rigidity, and muscular Irish JD. 2010. The mean measure of divergence:
entheses. American Journal of Physical its utility in model-free and model-bound
Anthropology 150:609–617. analyses relative to the Mahalanobis D2
Ichikawa M. 1981. Ecological and sociological distance for nonmetric traits. American Journal
importance of honey to the Mbuti net hunters, of Human Biology 22:378–395.
eastern Zaire. African Study Monographs Irish JD. 2012. Population continuity after all?
1:55–68. Potential late Pleistocene dental ancestors of
Infante P. 1974. Enamel hypoplasia in Apache Holocene Nubians have been found! American
Indian children. Ecology of Food and Nutrition Journal of Physical Anthropology Supplement
2:155–164. 54:172–173.
Infante P, Gillespie GM. 1974. An epidemiologic Irish JD. 2013. Afridonty: the “Sub-Saharan
study of linear enamel hypoplasia of deciduous African Dental Complex” revisited. In: Scott
anterior teeth in Guatemalan children. GR, Irish JD, editors. Anthropological
Archives of Oral Biology 19:1055–1061. Perspectives on Tooth Morphology: Genetics,
Infante P, Gillespie GM. 1977. Enamel hypoplasia Evolution, Variation. Cambridge, UK:
in relation to caries in Guatemalan children. Cambridge University Press; p. 278–295.
Journal of Dental Research 56:493–498. Irish JD, Davis SD, Lively RA. 1993.
Ingelmark BE. 1939. The skeletons. In: A bioarchaeological study of prehistoric
Thordeman B, editor. Armour from the Battle human, faunal, and cultural remains from
of Wisby 1361. Stockholm, Sweden: Kungl. Wilson Cove, Admiralty Island, Alaska. Arctic
Vitterhets Histoire och Antikvitets Akademien; Anthropology 30:103–119.
p. 149–209. Irish JD, Konigsberg L. 2007. The ancient
Inhorn MC, Brown PJ. 1990. The anthropology of inhabitants of Jebel Moya redux: measures of
infectious disease. Annual Review of population affinity based on dental
Anthropology 19:89–117. morphology. International Journal of
Inoue N, Ito G, Kamegai T. 1986. Dental pathology Osteoarchaeology 17:138–156.
of hunter-gatherers and early farmers in Irish JD, Turner CG II. 1987. More lingual surface
prehistoric Japan. In: Akazawa T, Aikens CM, attrition of the maxillary anterior teeth in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 495

American Indians: prehistoric Panamanians. high-seafoods diet and demographic stress.


American Journal of Physical Anthropology International Journal of Osteoarchaeology
73:209–213. 6:346–381.
Irish JD, Turner CG II, 1990. West African dental Ives R, Brickley M. 2005. Metacarpal
affinity of Late Pleistocene Nubians: peopling radiogrammetry: a useful indicator of bone loss
of the Eurafrican-South Asian triangle II. throughout the skeleton? Journal of
Homo 41:42–53. Archaeological Science 32:1552–1559.
Irish JD, Turner CG II, 1997. First evidence of Jackes J. 1994. Birth rates and bones. In: Herring
LSAMAT in non-native Americans: historic A, Chan L, editors. Strength in Diversity:
Senegalese from West Africa. American Journal A Reader in Physical Anthropology. Toronto,
of Physical Anthropology 102:141–146. ON: Canadian Scholars’ Press; p. 157–185.
Irish JJ. 1997. Characteristic high- and low- Jackes M. 2011. Representativeness and bias in
frequency dental traits in Sub-Saharan African archaeological skeletal samples. In: Agarwal
populations. American Journal of Physical SC, Glencross BA, editors. Social
Anthropology 102:455–467. Bioarchaeology. Chichester, UK: Wiley-
Irwin C, Mullally B, Ziada H, Allen E. 2008. Blackwell; p. 107–146.
Periodontics: 9. Periodontitis and systemic Jackes M, Meiklejohn C. 2008. The
conditions: is there a link? Dental Update paleodemography of central Portugal and the
35:92–101. Mesolithic-Neolithic transition. In: Bocquet-
İşcan MY, Miller-Shavitz P. 1985. Prehistoric Appel J-P, editor. Recent Advances in
syphilis in Florida. Florida Medical Annals Paleodemography: Data, Techniques, Patterns.
72:109–113. New York, NY: Springer; p. 209–258.
Israel H. 1968. Continuing growth in the human Jackson DW, Wiltse LL, Civicione RJ. 1976.
cranial skeleton. Archives of Oral Biology Spondylolysis in the female gymnasts. Clinical
13:133–138. Orthopaedics 117:68–73.
Israel H. 1973. Age factor and the pattern of Jacobi KP. 1997. Dental genetic structuring of a
change in craniofacial structures. American colonial Maya cemetery, Tipu, Belize. In:
Journal of Physical Anthropology 39:111–128. Whittington SL, Reed DM, editors. Bones of the
Israel H. 1977. The dichotomous pattern of Maya: Studies of Ancient Skeletons.
craniofacial expansion during aging. American Washington, DC: Smithsonian Institution
Journal of Physical Anthropology 47:47–52. Press; p. 138–153.
Issa SN, Sharma L. 2006. Epidemiology of Jacobi KP. 2000. Last Rites for the Tipu Maya:
osteoarthritis: an update. Current Genetic Structuring in a Colonial Cemetery.
Rheumatology Reports 8:7–15. Tuscaloosa, AL: University of Alabama Press.
Ito G, Shiono K, Inuzuka K, Hanihara K. 1983. Jacobi KP. 2007. Disabling the dead: human
Secular changes of tooth to denture base trophy taking in the prehistoric Southeast. In:
discrepancy during Japanese prehistoric and Chacon RJ, Dye DH, editors. The Taking and
historic ages. Journal of the Anthropological Displaying of Human Body Parts as Trophies
Society of Nippon 91:39–48. by Amerindians. New York, NY: Spring
Ivanhoe F. 1995. Secular decline in cranioskeletal Science + Business Media LLC; p. 299–338.
size over two millennia of interior central Jacobi KP, Cook DC, Corruccini RS, Handler JS.
California prehistory: relation to calcium 1992. Congenital syphilis in the past: slaves at
deficit in the reconstructed diet and Newton Plantation, Barbados, West Indies.
demographic stress. International Journal of American Journal of Physical Anthropology
Osteoarchaeology 5:213–253. 89:145–158.
Ivanhoe F, Chu PW. 1996. Cranioskeletal size Jacobs K. 1993. Human postcranial variation in
variation in San Francisco Bay prehistory: the Ukrainian Mesolithic-Neolithic. Current
relation to calcium deficit in the reconstructed Anthropology 34:311–324.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
496 References

Jacobs K. 1994. Human dento-gnathic metric Jantz RL, Owsley DW. 1984a. Long bone growth
variation in Mesolithic/Neolithic Ukraine: variation among Arikara skeletal populations.
possible evidence of demic diffusion in the American Journal of Physical Anthropology
Dnieper Rapids region. American Journal of 63:13–20.
Physical Anthropology 95:1–26. Jantz RL, Owsley DW. 1984b. Temporal changes in
Jacobs K. 1995. Returning to Oleni’ ostrov: social, limb proportionality among skeletal samples of
economic, and skeletal dimensions of a boreal Arikara Indians. Annals of Human Biology
forest Mesolithic cemetery. Journal of 11:157–164.
Anthropological Archaeology 14:359–403. Jantz RL, Owsley DW. 1994a. Growth and dental
Jankauskas R. 1999. Tuberculosis in Lithuania: development in Arikara children. In: Owsley
paleopathological and historical correlations. DW, Jantz RL, editors. Skeletal Biology in the
In: Pálfi G, Dutour O, Deak J, Hutas I, editors. Great Plains: Migration, Warfare, Health, and
Tuberculosis: Past and Present. Budapest/ Subsistence. Washington, DC: Smithsonian
Szeged, Hungary: Golden Book Publishers and Institution Press; p. 247–258.
Tuberculosis Foundation; p. 551–558. Jantz RL, Owsley DW. 1994b. White traders in the
Jantz RL. 1972. Cranial variation and Upper Missouri: evidence from the Swan Creek
microevolution in Arikara skeletal populations. site. In: Owsley DW, Jantz RL, editors. Skeletal
Plains Anthropologist 17:20–35. Biology in the Great Plains: Migration,
Jantz RL. 1973. Microevolutionary change in Warfare, Health, and Subsistence. Washington,
Arikara crania: a multivariate analysis. DC: Smithsonian Institution Press; p. 189–201.
American Journal of Physical Anthropology Jantz RL, Owsley DW. 2001. Variation among
38:15–26. early North American crania. American
Jantz RL. 1974. The Redbird Focus: cranial Journal of Physical Anthropology
evidence in tribal identification. Plains 114:146–155.
Anthropologist 19:5–13. Jantz RL, Owsley DW, Willey P. 1981.
Jantz RL. 1977. Craniometric relationships of Craniometric variation in the Northern and
Plains populations: historical and evolutionary Central Plains. In: Jantz RL, Ubelaker DH,
implications. Plains Anthropologist Memoir, editors. Progress in Skeletal Biology of Plains
No.13; p. 162–176. Populations. Plains Anthropologist Memoir,
Jantz RL. 1994. The social, historical, and No. 17; p. 19–29.
functional dimensions of skeletal variation. In: Jasienska G. 2010. Why women differ in ovarian
Owsley DW, Jantz RL, editors. Skeletal Biology function: genetic polymorphism,
in the Great Plains: Migration, Warfare, developmental conditions, and adult lifestyle.
Health, and Subsistence. Washington, DC: In: Muehlenbein MP, editor. Human
Smithsonian Institution Press; p. 175–178. Evolutionary Biology. Cambridge, UK:
Jantz RL. 2001. Cranial change in Americans: 1850– Cambridge University Press; p. 322–337.
1975. Journal of Forensic Science 46:784–787. Jaworowski I. 1968. Stable lead in fossil ice and
Jantz RL, Hunt DR, Falsetti AB, Key PJ. 1992. bones. Nature 217:152–153.
Variation among North Amerindians: analysis Jenkins CL. 1982. Factors in the aetiology of poor
of Boas’s anthropometric data. Human Biology growth in Belize. Cajanus 15:172–184.
64:435–461. Jenkins DP, Cochran TH. 1969. Osteoporosis: the
Jantz RL, Logan MH. 2010. Why does head form dramatic effect of disuse of an extremity.
change in children of immigrants? A reappraisal. Clinical Orthopaedics and Related Research
American Journal of Human Biology 22: 64:128–134.
702–707. Jim S, Ambrose SH, Evershed RP. 2004. Stable
Jantz RL, Meadows Jantz L. 2000. Secular change carbon isotopic evidence for differences in the
in craniofacial morphology. American Journal dietary origin of bone cholesterol, collagen and
of Human Biology 12:327–338. apatite: implications for their use in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 497

palaeodietary reconstruction. Geochimica et Johnston FE, Wainer H, Thissen D, MacVean RB.


Cosmochimica Acta 68:61–72. 1976. Hereditary and environmental
Jiménez SB. 1994. Occupational hazards in determinants of growth in height in a
19th-century Upper Canada. In: Herring A, longitudinal sample of children and youth of
Chan L, editors. Strength in Diversity: A Reader Guatemalan and European ancestry. American
in Physical Anthropology. Toronto, ON: Journal of Physical Anthropology 44:469–476.
Canadian Scholars’ Press; p. 345–364. Jones BC, Hann J, Scarry JF (editors). 1991. San
Jiménez-Brobeil SA, du Souich P, Al Oumaoui I. Pedro y San Pablo de Patale: A Seventeenth-
2009. Possible relationship of cranial traumatic Century Spanish Mission in Leon County,
injuries with violence in the south-east Iberian Florida. Florida Archaeology, No. 5.
peninsula from the Neolithic to the Bronze Age. Jones BC, Storey R, Widmer RJ. 1991. The Patale
American Journal of Physical Anthropology cemetery: evidence concerning the Apalachee
140:465–475. mission mortuary complex. In: Jones BC, Hann
Johannsdottir B, Thorarinsson F, Thordarson A, J, Scarry JF, editors. San Pedro y San Pablo de
Mangusson TE. 2005. Heritability of Patale: A Seventeenth-Century Spanish
craniofacial characteristics between parents Mission in Leon County, Florida. Florida
and offspring estimated from lateral Archaeology, No. 5; p. 109–136.
cephalograms. American Journal of Jones FW. 1908. The examination of the bodies of
Orthodontics and Dentofacial Orthopedics 100 men executed in Nubia in Roman times.
127:200–207. Britain Medical Journal 736–741.
Johansson SR, Horowitz S. 1986. Estimating Jones FW. 1910. General pathology (including
mortality in skeletal populations: influence of diseases of the teeth). In: Smith GE, Jones FW,
the growth rate on the interpretation of levels editors. Report on the Human Remains II. Cairo,
and trends during the transition to agriculture. Egypt: National Printing Company;
American Journal of Physical Anthropology p. 263–291.
71:233–250. Jones G. 1984. A History of the Vikings, Second
Johnson NW, Glick M, Mbuguye TN. 2006. Oral Edition. Oxford, UK: Oxford University Press.
health and general health. Advances in Dental Jones GD. 1978. The ethnohistory of the Guale
Research 19:118–121. coast through 1684. In: Thomas DH, Jones GD,
Johnston FE. 1962. Growth of the long bones of Durham RS, Larsen CS, editors. The
infants and young children at Indian Knoll. Anthropology of St. Catherines Island: 1.
American Journal of Physical Anthropology Natural and Cultural History. Anthropological
20:249–254. Papers of the American Museum of Natural
Johnston FE. 1969. Approaches to the study of History, No. 55; p. 178–210.
developmental variability in human skeletal Jones HH, Priest JD, Hayes WC, Techenor CC,
populations. American Journal of Physical Nagel DA. 1977. Humeral hypertrophy in
Anthropology 31:335–341. response to exercise. Journal of Bone and Joint
Johnston FE, Borden M, MacVean RB. 1975. The Surgery 59-A:204–208.
effects of genetic and environmental factors upon Jones J. 1876. Explorations of the Aboriginal
the growth of children in Guatemala City. In: Remains of Tennessee. Smithsonian
Watts ES, Johnston FE, Lasker GW, editors. Contributions to Knowledge, 259.
Biosocial Interrelations in Population Adaptation. Jones S, Quinn RL. 2009. Prehistoric Fijian diet
The Hague, the Netherlands: Mouton; p. 377–388. and subsistence: integration of faunal,
Johnston FE, Snow CE. 1961. The reassessment of ethnographic, and stable isotope evidence from
the age and sex of the Indian Knoll skeletal the Lau Island group. Journal of Archaeological
population: demographic and methodological Science 36:2742–2754.
aspects. American Journal of Physical Jonsson R, Howland BE, Bowden GH. 1988.
Anthropology 19:237–244. Relationships between periodontal health,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
498 References

salivary steroids, and Bacteroides intermedius American Journal of Physical Anthropology


in males, pregnant and non-pregnant women. 131:324–333.
Journal of Dental Research 67:1062–1069. Judd MA. 2008. The parry problem. Journal of
Jordan JM. 2000. Systemic risk factors for Archaeological Science 35:1658–1666.
osteoarthritis. Annals of Internal Medicine Judd MA. 2012. Gabati: A Meroitic, Post-Meroitic,
133:637–643. and Medieval Cemetery in Central Sudan.
Jordan JM, Helmick CG, Renner JB, et al. 2007. Volume 2. The Physical Anthropology. British
Prevalence of knee symptoms and radiographic Archaeological Reports, International Series,
and symptomatic knee osteoarthritis in African No. 2442.
Americans and Caucasians: The Johnston Judd M, Irish J. 2009. Dying to serve: the mass
County Osteoarthritis Project. Journal of burials at Kerma. Antiquity 83:1–14.
Rheumatology 34:172–180. Judd MA, Redfern R. 2012. Trauma. In: Grauer AL,
Jordan JM, Linder GF, Fryer JG, Renner JB. 1995. editor. A Companion to Paleopathology.
The impact of arthritis in rural populations. Chichester, UK: Wiley-Blackwell; p. 359–379.
Arthritis Care and Research 8:242–250. Judd MA, Roberts CA. 1998. Fracture patterns at
Jordan JM, Syddall H, Dennison EM, Cooper C, the medieval leper hospital in Chichester.
Arden NK. 2005. Birthweight, vitamin D American Journal of Physical Anthropology
receptor gene polymorphism, and risk of 105:43–56.
lumbar spinal osteoarthritis. Journal of Judd MA, Roberts CA. 1999. Fracture trauma in a
Rheumatology 32:678–683. medieval British farming village. American
Jordan RA, Lucaciu A, Fotouhi K, et al. 2011. Pilot Journal of Physical Anthropology 109:229–244.
pathfinder survey of oral hygiene and Jungers WL, Minns RJ. 1979. Computed
periodontal conditions in the rural population tomography and biomechanical analysis of
of The Gambia (West Africa). International fossil long bones. American Journal of Physical
Journal of Dental Hygiene 9:53–59. Anthropology 50:285–290.
Jørkov MLS, Jørgensen L, Lynnerup N. 2010. Jurmain R. 1990. Paleoepidemiology of a central
Uniform diet in a diverse society. Revealing California prehistoric population from CA-
new dietary evidence of the Danish Roman Iron ALA-329: II. Degenerative disease. American
Age based on stable isotope analysis. American Journal of Physical Anthropology 83:83–94.
Journal of Physical Anthropology Jurmain R. 1999. Stories from the Skeleton:
143:523–533. Behavioral Reconstruction in Human
Joshipura KJ, Rimm EB, Douglass CW, et al. 1996. Osteology. Amsterdam, the Netherlands:
Poor oral health and coronary heart disease. Gordon and Breach Publishers.
Journal of Dental Research 75:1631–1636. Jurmain RD. 1977a. Stress and etiology of
Jovanovic L. 2000. A tincture of time does not osteoarthritis. American Journal of Physical
turn the tide: type 2 diabetes trends in Anthropology 46:353–366.
offspring of type 2 diabetic mothers. Diabetes Jurmain RD. 1977b. Paleoepidemiology of
Care 23:1219–1220. degenerative knee disease. Medical
Judd M. 2002. Ancient injury recidivism: an Anthropology 1:1–14.
example from the Kerma period of ancient Jurmain RD. 1978. Paleoepidemiology of
Nubia. International Journal of degenerative joint disease. Medical College of
Osteoarchaeology 12:89–106. Virginia Quarterly 14:45–46.
Judd M. 2004. Trauma in the city of Kerma: Jurmain RD. 1980. The pattern of involvement of
ancient versus modern injury patterns. appendicular degenerative joint disease.
International Journal of Osteoarchaeology American Journal of Physical Anthropology
14:34–51. 53:143–150.
Judd MA. 2006. Continuity of interpersonal Jurmain RD. 1991. Paleoepidemiology of trauma
violence between Nubian communities. in a prehistoric central California population.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 499

In: Ortner DJ, Aufderheide AC, editors. Human American Journal of Physical Anthropology
Paleopathology: Current Syntheses and Future 109:485–500.
Options. Washington, DC: Smithsonian Kaifu Y. 2000. Tooth wear and compensatory
Institution Press; p. 241–248. modification of the anterior dentoalveolar
Kacki S, Rahalison L, Rajerison M, Ferroglio E, complex in humans. American Journal of
Bianucci R. 2011. Black Death in the rural Physical Anthropology 111:369–392.
cemetery of Saint-Laurent-de-la-Cabrerisse Kakaliouras AM. 2008. Leaving few bones
Audi-Languedoc, southern France, 14th unturned: recent work on repatriation by
century: immunological evidence. Journal of osteologists. American Anthropologist
Archaeological Science 38:581–587. 110:44–52.
Kaestle F. 2010. Paleogenetics: ancient DNA in Kakaliouras AM. 2012. An anthropology of
anthropology. In: Larsen CS, editor. repatriation: contemporary physical
A Companion to Biological Anthropology. anthropological and Native American
Malden, MA: Wiley-Blackwell; p. 427–441. ontologies of practice. Current Anthropology
Kaestle FA. 1995. Mitochondrial DNA evidence for 53(S5):S210–S221.
the identity of the descendants of the Kamegai T, Kuragano S, Hanihara K. 1982. Secular
prehistoric Stillwater Marsh populations. In: change of dentofacial morphology during
Larsen CS, Kelly RL, editors. Bioarchaeology of Japanese historic ages. Journal of the
the Stillwater Marsh: Prehistoric Human Anthropological Society of Nippon
Adaptation in the Western Great Basin. 90:303–314.
Anthropological Papers of the American Kannegaard Nielsen E, Brinch Petersen E. 1993.
Museum of Natural History, No. 77; p. 73–80. Burials, people and dogs. In: Hvass S,
Kaestle FA, Lorenz JG, Smith DG. 1999. Molecular Storgaard B, editors. Digging into the Past:
genetics and the Numic expansion: a molecular 25 Years of Archaeology in Denmark.
investigation of the prehistoric inhabitants of Copenhagen: Royal Society of Antiquaries;
Stillwater Marsh. In: Hemphill BE, Larsen CS, p. 76–81.
editors. Prehistoric Lifeways in the Great Basin Kannus P, Haapasalo H, Sankelo M et al. 1995.
Wetlands: Bioarchaeological Reconstruction Effect of starting age of physical activity on
and Interpretation. Salt Lake City, UT: bone mass in the dominant arm of tennis and
University of Utah Press; p. 167–183. squash players. Annals of Internal Medicine
Kaestle FA, Smith DG. 2001. Ancient 123:27–31.
mitochondrial DNA evidence for prehistoric Kanz F, Grossschmidt K. 2006. Head injuries of
population movement: the Numic expansion. Roman gladiators. Forensic Science
American Journal of Physical Anthropology International 160:207–216.
115:1–12. Kapferer B. 2012. Legends of People, Myths of
Kaidonis JA, Richards LC, Townsend GC. 1993. State: Violence, Intolerance, and Political
Nature and frequency of dental wear facets in Culture in Sri Lanka and Australia. New York,
an Australian Aboriginal population. Journal NY: Berghahn Books.
of Oral Rehabilitation 20:333–340. Kaplan RF, Dorn BC, Hood JS, McGann PD. 1977.
Kaifu Y. 1997. Changes in mandibular Musculoskeletal system. In: Tedeschi CG,
morphology from the Jomon to modern periods Eckert WG, Tedeschi LG, editors. Forensic
in eastern Japan. American Journal of Physical Medicine: A Study in Trauma and
Anthropology 104:227–243. Environmental Hazards, Volume 1. Mechanical
Kaifu Y. 1998. Sex differences in tooth wear in the Trauma. Philadelphia, PA: W.B. Saunders;
Japanese. Bulletin of the National Science p. 259–361.
Museum, Tokyo Series D 24:49–59. Kasai K, Kawamura A. 2001. Correlation between
Kaifu Y. 1999. Changes in pattern of tooth wear buccolingual inclination and wear of
from prehistoric to recent periods in Japan. mandibular teeth in ancient and modern

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
500 References

Japanese. Archives of Oral Biology Katzenberg MA. 2008. Stable isotope analysis: a
46:269–273. tool for studying past diet, demography, and
Katz SH, Hediger ML, Valleroy LA. 1974. life history. In: Katzenberg MA, Saunders SR,
Traditional maize processing techniques in the editors. Biological Anthropology of the Human
New World. Science 184:765–773. Skeleton, Second Edition. Hoboken, NJ: Wiley-
Katzenberg MA. 1977. An investigation of spinal Liss; p. 413–442.
disease in a Midwest aboriginal population. Katzenberg MA, Harrison RG. 1997. What’s in a
Yearbook of Physical Anthropology bone? Recent advances in archaeological bone
20:349–355. chemistry. Journal of Archaeological Research
Katzenberg MA. 1984. Chemical Analysis of 5:265–293.
Prehistoric Human Bone from Five Temporally Katzenberg MA, Herring DA, Saunders SR. 1996.
Distinct Populations in Southern Ontario. Weaning and infant mortality: evaluating the
Archaeological Survey of Canada, Mercury skeletal evidence. Yearbook of Physical
Series Paper, No. 129. Anthropology 39:177–199.
Katzenberg MA. 1991a. Stable isotope analysis of Katzenberg MA, McKenzie HG, Losey RJ,
remains from the Harvie family. In: Saunders Goriunova OI, Weber A. 2012. Prehistoric
S, Lazenby R, editors. The Links that Bind: The dietary adaptations among hunter-fisher-
Harvie Family Nineteenth-Century Burying gatherers from the Little Sea of Lake Baikal,
Ground. Occasional Papers in Northeastern Siberia, Russian Federation. Journal of
Archaeology, 5; p. 65–69. Archaeological Science 39:2612–2626.
Katzenberg MA. 1991b. Analysis of stable isotopes Katzenberg, MA, Saunders SR, editors. 2008.
of carbon and nitrogen. In: Pfeiffer S, Biological Anthropology of the Human
Williamson RF, editors. Snake Hill: An Skeleton, Second Edition. Hoboken, NJ: Wiley-
Investigation of a Military Cemetery from the Liss.
War of 1812. Toronto, ON: Dundurn Press; Katzenberg MA, Schwarcz HP, Knyf M, Melbye FJ.
p. 247–255. 1995. Stable isotope evidence for maize
Katzenberg MA. 1992. Changing diet and health in horticulture and paleodiet in southern Ontario,
pre- and protohistoric Ontario. In: Huss- Canada. American Antiquity 60:335–350.
Ashmore R, Schall J, Hediger M, editors. Health Katzenberg MA, Weber A. 2009. Stable isotope
and Lifestyle Change. MASCA Research ecology and paleodiet in the Lake Baikal region
Papers in Science and Archaeology, No. 9; of Siberia. Journal of Archaeological Science
p. 23–31. 26:651–659.
Katzenberg MA. 1993. Applications of elemental Kay RF. 1987. Analysis of primate dental
and isotopic analysis to problems in Ontario microwear using image processing techniques.
prehistory. In: Sandford MK, editor. Scanning Microscopy 1:657–662.
Investigations of Ancient Human Tissue: Kazarian LE, Von Gierke HE. 1969. Bone loss as a
Chemical Analyses in Anthropology. result of immobilization and chelation:
Langhorne, PA: Gordon and Breach Scientific preliminary results in Macaca mulatta. Clinical
Publishers; p. 335–360. Orthopaedics and Related Research 65:67–75.
Katzenberg MA. 2006. Prehistoric maize in Keegan WF, DeNiro MJ. 1988. Stable carbon- and
southern Ontario: contributions from stable nitrogen-isotope ratios of bone collagen used
isotope studies. In: Staller JJ, Staller R, Tykot R, to study coral-reef and terrestrial components
Benz B, editors. Histories of Maize: of prehistoric Bahamian diet. American
Multidisciplinary Approaches to the Antiquity 53:320–336.
Prehistory, Linguistics, Biogeography, Keeley LH. 1996. War Before Civilization.
Domestication, and Evolution of Maize. New York, NY: Oxford University Press.
Amsterdam, the Netherlands: Elsevier; Keener CS. 1999. An ethnohistorical analysis of
p. 263–273. Iroquois assault tactics used against fortified

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 501

settlements of the northeast in the 17th Kellner CM, Schoeninger MJ. 2007. A simple
century. Ethnohistory 46:777–807. carbon isotope model for reconstructing
Keenleyside A. 1998. Skeletal evidence of health prehistoric human diet. American Journal of
and disease in pre-contact Alaskan Eskimos Physical Anthropology 133:1112–1127.
and Aleuts. American Journal of Physical Kellner CM, Schoeninger MJ. 2008. Wari’s
Anthropology 107:51–70. imperial influence on local Nasca diet: the
Keenleyside A. 2006. Skeletal biology: arctic and stable isotope evidence. Journal of
subarctic. In: Ubelaker DH, editor. Handbook of Anthropological Archaeology 27:226–243.
North American Indians: 3. Environment, Kelly JA, Tykot RH, Milanich JT. 2006. Evidence for
Origins, and Population. Washington, DC: early use of maize in peninsular Florida. In:
Smithsonian Institution Press; p. 524–531. Staller J, editor. Histories of Maize. Amsterdam,
Keenleyside A, Song X, Chettle DR, Webber CE. the Netherlands: Elsevier; p 249–361.
1996. The lead content of human bones from Kelly RL. 1992. Mobility/sedentism: concepts,
the 1845 Franklin Expedition. Journal of archaeological measures, and effects. Annual
Archaeological Science 23:461–465. Review of Anthropology 21:43–66.
Keil EG. 1933. The importance of nutrition in the Kelly RL. 2001. Prehistory of the Carson Desert
prevention and cure of leprosy. International and Stillwater Mountains: Environment,
Journal of Leprosy 1:393–397. Mobility, and Subsistence in a Great Basin
Keith A. 1916. Is the British facial form changing? Wetland. University of Utah Anthropological
Nature 98:198. Papers, No. 123.
Keith A. 1924. Concerning Certain Structural Kelly RL. 2013. The Lifeways of Hunters –
Changes which are Taking Place in Our Jaws Gatherers: The Foraging Spectrum. New York,
and Teeth. London, UK: Dental Board of United NY: Cambridge University Press.
Kingdom. Kemkes-Grottenthaler A. 2005. The short die
Keith A. 1950. An Autobiography. London, UK: young: the interrelationship between stature
Watts & Company. and longevity – evidence from skeletal
Keith MS, Armelagos GJ. 1988. An example of in remains. American Journal of Physical
vivo tetracycline labelling: reply to Anthropology 128:340–347.
Piepenbrink. Journal of Archaeological Science Kemp BM, Malhi RS, McDonough J, et al. 2007.
15:595–601. Genetic analysis of early Holocene skeletal
Kelley JO, Angel JL. 1987. Life stresses in slavery. remains from Alaska and its implications for
American Journal of Physical Anthropology the settlement of the Americas. American
74:199–211. Journal of Physical Anthropology
Kelley M, Micozzi M. 1984. Rib lesions in 132:605–621.
pulmonary tuberculosis. American Journal of Kemp BM, Tung TA, Summar ML. 2009. Genetic
Physical Anthropology 65:381–386. continuity after the collapse of the Wari
Kellgren JH, Lawrence JS. 1958. Osteo-arthrosis Empire: mitochondrial DNA profiles from Wari
and disc degeneration in an urban population. and post-Wari populations in the ancient
Annals of the Rheumatic Diseases Andes. American Journal of Physical
17:388–397. Anthropology 140:80–91.
Kellner CM, Schoeninger M, Spielmann KA, Moore Kennedy B. 1981. Marriage Patterns in an Archaic
K. 2010. Stable isotope data show temporal Population: A Study of Skeletal Remains from
stability in diet at Pecos Pueblo and diet Port au Choix, Newfoundland. Archaeological
variation among Southwest Pueblos. In: Survey of Canada, Mercury Series Paper,
Morgan ME, editor. Pecos Pueblo Revisited: The No. 104.
Biological and Social Context. Harvard Kennedy BV. 1989. Variation in δ13C Values of
University Papers of the Peabody Museum of Post-Medieval Europeans. PhD Dissertation,
Archaeology and Ethnology, 85; p. 79–91. University of Calgary, AB.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
502 References

Kennedy KAR. 1983. Morphological variations in craniometric evidence. In: Owsley DW, Jantz
ulnar supinator crests and fossae as identifying RL, editors. Skeletal Biology in the Great
markers of occupational stress. Journal of Plains: Migration, Warfare, Heath, and
Forensic Sciences 28:871–876. Subsistence. Washington, DC: Smithsonian
Kennedy KAR. 1984. Growth, nutrition, and Institution Press; p. 179–187.
pathology in changing paleodemographic Key PJ, Jantz RL. 1990. Statistical assessment of
settings in South Asia. In: Cohen MN, population variability: a methodological
Armelagos GJ, editors. Paleopathology at the approach. American Journal of Physical
Origins of Agriculture. Orlando, FL: Academic Anthropology 82:53–59.
Press; p. 169–192. Keyser-Tracqui C, Crubézy E, Ludes B. 2003.
Kennedy KAR. 1989. Skeletal markers of Nuclear and mitochondrial analysis of a 2000-
occupational stress. In: İşcan MY, Kennedy year-old necropolis in the Egyin Gol Valley
KAR, editors. Reconstruction of Life from the (Mongolia). American Journal of Human
Skeleton. New York, NY: Alan R. Liss; Genetics 73:247–260.
p. 129–160. Keyserling WM. 1988. CTD statistics: prevalence,
Kennedy KAR. 1994. Identification of sacrificial incidence and severity. In: Putz-Anderson,
and massacre victims in archaeological sites: editor. Cumulative Trauma Disorders:
the skeletal evidence. Man and Environment A Manual for Musculoskeletal Diseases of the
19:247–251. Upper Limbs. London, UK: Taylor & Francis;
Kennedy KAR. 1998. Markers of occupational p. 129–141.
stress: conspectus and prognosis of research. Kidder AV. 1932. The Artifacts of Pecos. Papers of
International Journal of Osteoarchaeology the Southwest Expedition 6, Department of
8:305–310. Archaeology, Phillips Academy, Andover, UK.
Kennett DJ. 2005. The Island Chumash: Kieser JA. 1990. Human Adult Odontometrics: The
Behavioral Ecology of a Maritime Society. Study of Variation in Adult Tooth Size.
Berkeley, CA: University of California Press. Cambridge, UK: Cambridge University Press.
Kent S. 1986. The influence of sedentism and Kieser JA, Groeneveld HT. 1998. Fluctuating
aggregation on porotic hyperostosis and dental asymmetry and prenatal exposure to
anaemia: a case study. Man 21:605–636. tobacco smoke. In: Lukacs JR, editor. Human
Kerr NW. 1998. The prevalence and natural history Dental Development, Morphology, and
of periodontal disease in Britain from Pathology. University of Oregon
prehistoric to modern times. British Dental Anthropological Papers, No. 54; p. 287–297.
Journal 185:527–535. Kieser JA, Groeneveld HT, Preston CB. 1985.
Kerr RA. 1991. Dinosaurs and friend snuffed out? Craniofacial correlates of attrition. Angle
Science 251:160–162. Orthodontist 55:329–335.
Kestle S. 1988. Subsistence and sex roles. In: Kieser JA, Groeneveld HT, Preston CB. 1986a.
Blakely RL, editor. The King Site: Continuity and Fluctuating dental asymmetry as a measure of
Contact in Sixteenth Century Georgia. Athens, odontogenic canalization in man. American
GA: University of Georgia Press.; p. 63–72. Journal of Physical Anthropology 71:437–444.
Keusch GT, Farthing MJG. 1986. Nutrition and Kieser JA, Groeneveld HT, Preston CB. 1986b.
infection. Annual Review of Nutrition Fluctuating odontometric asymmetry in the
6:131–154. Lengua Indians of Paraguay. Annals of Human
Key PJ. 1983. Craniometric Relationships among Biology 13:489–498.
Plains Indians. University of Tennessee, Kilgore L, Jurmain R, Van Gerven D. 1997.
Department of Anthropology, Reports of Palaeoepidemiological patterns of trauma in a
Investigations, No. 34. medieval Nubian skeletal population.
Key PJ. 1994. Relationships of the Woodland International Journal of Osteoarchaeology
period on the Northern and Central Plains: the 7:103–114.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 503

Kiliaridis S. 1995. Masticatory muscle influence on Kiratli BJ. 1992. Bone mineral content and body
craniofacial growth. Acta Odontologica composition responses to unilateral disuse
Scandinavica 53:196–202. (hemiplegia). American Journal of Physical
Killgrove K. 2010. Identifying immigrants in Anthropology S14:100–101.
Imperial Rome using strontium isotope Kirch PV. 1984. The Evolution of Polynesian
analysis. Journal of Roman Archaeology Chiefdoms. Cambridge, UK: Cambridge
S78:157–174. University Press.
Killgrove K, Tykot RH. 2013. Food for Rome: a Kjellström A. 2012. Possible cases of leprosy and
stable isotope investigation of diet in the tuberculosis in Medieval Sigtuna, Sweden.
Imperial period (1st–3rd centuries AD). Journal International Journal of Osteoarchaeology
of Anthropological Archaeology 32:28–38. 22:261–283.
Kimura K. 1984. Studies on growth and Kjellström A, Stora J, Possnert G, Linderholm A.
development in Japan. Yearbook of Physical 2009. Dietary patterns and social structures in
Anthropology 27:179–214. Medieval Sigtuna, Sweden, as reflected in
Kimura T. 1971. Cross-section of human lower stable isotope values in human skeletal
limb bones viewed from strength of materials. remains. Journal of Archaeological Science
Journal of the Anthropological Society of 36:2689–2699.
Nippon 79:323–336. Klaus HD. 2012. Bioarchaeology of structural
Kimura T. 1974. Mechanical characteristics of violence: theoretical model and case study. In:
human lower leg bones. Journal of the Faculty Martin DL, Harrod RP, Perez VR, editors. The
of Science, University of Tokyo Section 5 Bioarchaeology of Violence. Gainesville, FL:
4:319–393. University Press of Florida; p. 29–62.
Kimura T. 2002. Primate limb bones and Klaus HD. 2014a. A history of violence in the
locomotor types in arboreal or terrestrial Lambayeque Valley: conflict and death from
environments. Zeitschrift für Morphologie the late pre-Hispanic apogee to European
83:201–219. colonization of Peru (A.D. 900–1750). In:
Kimura T, Takahashi H. 1982. Mechanical Knüsel C, Smith MJ, editors. The Routledge
properties of cross section of lower limb bones Handbook of the Bioarchaeology of Human
in Jomon Man. Journal of the Anthropological Conflict. London, UK: Routledge; p. 389–414.
Society of Nippon 90(S):105–118. Klaus HD. 2014b. Subadult scurvy in Andean
King JA, Ubelaker DH. 1996. Living and Dying on the South America: evidence of vitamin C
17th Century Patuxent Frontier. Crownsville, deficiency in the late pre-Hispanic and
MD: Maryland Historical Trust Press. Colonial Lambayeque Valley, Peru.
King JW, Brelsford HJ, Tullos HS. 1969. Analysis International Journal of Paleopathology
of the pitching arm of the professional baseball 5:34–45.
pitcher. Clinical Orthopaedics 67:116–123. Klaus HD, Centurion J, Curo M. 2010.
King L, Harris EF, Tolley EA. 1993. Heritability of Bioarchaeology of human sacrifice: violence,
cephalometric and occlusal variables as identity and the evolution of ritual killing at
assessed from siblings with overt Cerro Cerrillos, Peru. Antiquity 84:1102–1122.
malocclusions. American Journal of Klaus HD, Larsen CS, Tam ME. 2009. Economic
Orthodontics and Dentofacial Orthopedics intensification and degenerative joint disease:
104:121–131. life and labor on the postcontact north coast of
King T, Humphrey LT, Hillson SW. 2005. Linear Peru. American Journal of Physical
enamel hypoplasias as indicators of systemic Anthropology 139:204–221.
physiological stress: evidence from two known Klaus HD, Tam ME. 2009. Contact in the Andes:
age-at-death and sex populations from bioarchaeology of systemic stress in colonial
postmedieval London. American Journal of Morrope, Peru. American Journal of Physical
Physical Anthropology 128:547–559. Anthropology 138:356–368.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
504 References

Klaus HD, Tam ME. 2010. Oral health and the Tiwanaku- and Chiribaya-affiliated sites in the
postcontact adaptive transition: a contextual Andes. American Journal of Physical
reconstruction of diet in Morrope, Peru. Anthropology 132:25–39.
American Journal of Physical Anthropology Knudson JK, Price TD, Buikstra JE, Blom DE. 2004.
141:594–609. The use of strontium isotope analysis to
Klaus HD, Toyne JM. 2015. Bodies and blood: investigate Tiwanaku migration and mortuary
Middle Sicán human sacrifice in the ritual in Bolivia and Peru. Archaeometry
Lambayeque Valley Complex (A.D. 900–1100). 46:5–18.
In: Klaus HD, Toyne JM, editors. Knudson KJ, O’Donnabhain B, Carver C, Cleland R,
Reconstructing Ritual Killing on the North Price TD. 2012. Migration and Viking Dublin:
Coast of Peru: Archaeological Studies of paleomobility and paleodiet through isotopic
Sacrifice and Violence in the Ancient Andes. analyses. Journal of Archaeological Science
Austin, TX: University of Texas Press, in press. 39:308–320.
Klaus HD, Wilbur AK, Temple DH, et al. 2010. Knudson KJ, Pestle WJ, Torres-Rouff C,
Tuberculosis on the north coast of Peru: Pimentel G. 2012. Assessing the life history of
skeletal and molecular paleopathology of late an Andean traveler through biogeochemistry:
pre-Hispanic and postcontact mycobacterial stable and radiogenic isotope analyses of
disease. Journal of Archaeological Science archaeological human remains from northern
37:2587–2597. Chile. International Journal of
Klenerman L, Swanson SAV, Freeman MAR. 1967. Osteoarchaeology 22:435–451.
A method for the clinical estimation of the Knudson KJ, Stojanowski CM. 2008. New
strength of a bone. Proceedings of the Royal directions in bioarchaeology: recent
Society of Medicine 60:10–14. contributions to the study of human social
Klingenberg CP. 2003. A developmental identities. Journal of Archaeological Research
perspective on developmental instability: 16:397–432.
theory, models, and mechanisms. In: Polak M, Knudson KJ, Stojanowski CM, editors. 2009.
editor. Developmental Instability: Causes and Bioarchaeology and Identity in the Americas.
Consequences. New York, NY: Oxford Tallahassee, FL: University Press of Florida.
University Press; p. 14–31. Knudson KJ, Torres-Rouff C. 2009. Investigating
Knapp M, Horsburgh KA, Prost S, et al. 2012. cultural heterogeneity in San Pedro de
Complete mitochondrial DNA genome Atacama, northern Chile, through
sequences from the first New Zealanders. biogeochemistry and bioarchaeology.
Proceedings of the National Academy of American Journal of Physical Anthropology
Sciences 109:18350–18354. 138:473–485.
Knauft BM. 1987. Reconsidering violence in Knudson KJ, Tung TA. 2011. Investigating
simple human societies: homicide among the regional mobility in the southern hinterland of
Gebusi of New Guinea. Current Anthropology the Wari empire. American Journal of Physical
28:457–500. Anthropology 145:299–310.
Knowles AK. 1983. Acute traumatic lesions. In: Knudson KJ, Williams HM, Buikstra JE, et al. 2010.
Hart GD, editor. Disease in Ancient Man. Introducing 88/86Sr analysis in archaeology: a
Agincourt, ON: Clarke Irwin; p. 61–83. demonstration of the utility of strontium
Knudson JK. 2008. Tiwanaku influence in the isotope fractionation in paleodietary studies.
south central Andes: strontium isotope Journal of Archaeological Science
analysis and Middle Horizon migration. Latin 37:2352–2364.
American Antiquity 19:3–23. Knüsel CJ. 2000. Activity-related changes in
Knudson JK, Price TD. 2007. Utility of multiple casualties from the medieval battle of Towton,
chemical techniques in archaeological A.D. 1461. In: Fiorato V, Boylton A, Knüsel CJ,
residential mobility studies: case studies from editors. Blood Red Roses: The Archaeology of a

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 505

Mass Grave from Towton, A.D. 1461. Oxford, Kolman CJ, Sambuughin N, Bermingham E. 1996.
UK: Oxbow Books; p. 103–118. Mitochondrial DNA analysis of Mongolian
Knüsel CJ. 2010. Bioarchaeology: a synthetic populations and implications for the origin
approach. Bulletins et Memoires de la Société of New World founders. Genetics
d’Anthropologie de Paris 22:62–73. 142:1321–1334.
Knüsel CJ. 2014. Courteous knights and cruel Kolodny Y, Luz B, Navon O. 1983. Oxygen isotope
avengers: a consideration of the changing variations in phosphate of biogenic apatites:
social context of medieval warfare from the 1. Fish bone apatite – rechecking the rules of
perspective of human remains. In: Knüsel C, the game. Earth and Planetary Science Letters
Smith MJ, editors. The Routledge Handbook of 64:398–404.
the Bioarchaeology of Human Conflict. London, Komlos J. 1987. The height and weight of West
UK: Routledge; p. 263–282. Point cadets: dietary change in antebellum
Knüsel CJ, Outram AK. 2006. Fragmentation of the America. Journal of Economic History
body: comestibles, compost, or customary rite? 47:897–927.
In: Goland R, Knüsel C, editors. Social Komlos J. 1989. Nutrition and Economic
Archaeology of Funerary Remains. Oxford, UK: Development in the Eighteenth-Century
Oxbow Books; p. 253–278. Habsburg Monarchy: An Anthropometric
Knüsel CJ, Roberts CA, Boylston A. 1996. When History. Princeton: Princeton University Press.
Adam delved. . .an activity-related lesion in Komlos J, editor. 1994. Stature, Living Standards,
three human skeletal populations. American and Economic Development: Essays in
Journal of Physical Anthropology Anthropometric History. Chicago: University of
100:427–434. Chicago Press.
Knüsel C., Smith MJ, editors. 2014. The Routledge Komlos J. 2009. Anthropometric history: an
Handbook of the Bioarchaeology of Human overview of a quarter century of research.
Conflict. Oxford, UK: Routledge. Anthropologischer Anzeiger 67:341–356.
Kobayashi K. 1967. Trend in the length of life Komlos J, Kriwy P. 2002. Social status and adult
based on human skeletons from prehistoric to heights in the two Germanies. Annals of
modern times in Japan. Journal of the Faculty Human Biology 29:641–648.
of Science, University of Tokyo 3:109–160. Konig KG. 1970. Feeding regimens and caries.
Kohler TA, Glaude MP, Bocquet-Appel J-P, Kemp Journal of Dental Research 49:1327–1333.
BM. 2008. The Neolithic demographic Konig KG, Larson RH, Guggenheim B. 1969.
transition in the U.S. Southwest. American A strain-specific eating pattern as a factor
Antiquity 73:645–669. limiting the transmissibility of caries activity in
Kohn LAP, Bennett KA. 1986. Fluctuating rats. Archives of Oral Biology 14:91–103.
asymmetry in fetuses of diabetic rhesus Konigsberg LW. 1987. Population Genetic Models
monkeys. American Journal of Physical for Interpreting Prehistoric Intra-Cemetery
Anthropology 71:477–483. Biological Variation. PhD Dissertation,
Kohrt WM, Bloomfield SA, Little KD, Nelson ME, Northwestern University, Evanston, IL.
Yingling VR. 2004. Physical activity and bone Konigsberg LW. 1988. Migration models of
health. Medicine and Science in Sports and prehistoric postmarital residence. American
Exercise 36:1985–1996. Journal of Physical Anthropology 77:471–482.
Kolata AL. 1993. The Tiwanaku: Portrait of an Konigsberg LW. 1990. Temporal aspects of
Andean Civilization. Oxford, UK: Blackwell. biological distance: serial correlation and trend
Kolman CJ, Centurion-Lara A, Lukehart SA, in a prehistoric skeletal lineage. American
Owsley DW, Tuross N. 1999. Identification of Journal of Physical Anthropology 82:45–52.
Treponema pallidum subspecies pallidum in a Konigsberg LW. 2006. A post-Neumann history of
200-year-old skeletal specimen. Journal of biological and genetic distance studies in
Infectious Diseases 108:2060–2063. bioarchaeology. In: Buikstra JE, Beck LA,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
506 References

editors. Bioarchaeology: The Contextual Kreshover SJ. 1944. The pathogenesis of enamel
Analysis of Human Remains. Burlington, MA: hypoplasia: an experimental study. Journal of
Academic Press; p. 263–280. Dental Research 23:231–239.
Konigsberg LW, Buikstra JE. 1995. Regional Kreshover SJ. 1960. Metabolic disturbances in
approaches to the investigation of past human tooth formation. Annals of the New York
biocultural structure. In: Beck LA, editor. Academy of Sciences 85:161–167.
Regional Approaches to Mortuary Practices. Kreshover SJ, Clough OW. 1953a. Prenatal
New York, NY: Plenum Press; p. 191–220. influences on tooth development: I. Alloxan
Konigsberg LW, Frankenburg SR. 1992. diabetes in rats. Journal of Dental Research
Estimation of age structure in anthropological 32:246–261.
demography. American Journal of Physical Kreshover SJ, Clough OW. 1953b. Prenatal
Anthropology 89:235–256. influences on tooth development: II.
Konigsberg LW, Frankenberg SR. 1994. Artificially induced fever in rats. Journal of
Paleodemography: “not quite dead.” Dental Research 32:565–572.
Evolutionary Anthropology 3:92–105. Kricun ME. 1994. Paleoradiology of the prehistoric
Konigsberg LW, Frankenberg SR. 2002. Australian Aborigines. American Journal of
Deconstructing death in paleodemography Roentgenology 163:241–247.
American Journal of Physical Anthropology Krigbaum J. 2007. Prehistoric dietary transitions in
117:297–307. tropical Southeast Asia: stable isotope and dental
Konigsberg LW, Owsley SD. 1995. Multivariate caries evidence from two sites in Malaysia. In:
quantitative genetics of anthropometric traits Cohen MN, Crane-Kramer GMM, editors. Ancient
from the Boas data. Human Biology Health: Skeletal Indicators of Agricultural and
67:481–498. Economic Intensification. Gainesville, FL:
Konomi N, Lebwohl E, Mowbray K, Tattersall I, University Press of Florida; p. 273–285.
Zhang D. 2002. Detection of mycobacterial Kroeber AL. 1925. Handbook of the Indians of
DNA in Andean mummies. Journal of Clinical California. New York, NY: Dover Publications.
Microbiology 40:4738–4740. Krueger HW, Sullivan CH. 1984. Models for
Koren O, Spor A, Felin J, et al. 2011. Human oral, carbon isotope fractionation between diet and
gut, and plaque microbiota in patients with bone. In: Turnlund JE, Johnson FE, editors.
atherosclerosis. Proceedings of the National Stable Isotopes in Nutrition. American
Academy of Sciences 108(S1):4592–4598. Chemical Society, Symposium Series 258;
Koritzer RT. 1968. Apparent tooth preparation in a p. 205–222.
Middle Mississippi Indian culture. Journal of Kuckelman KA, Lightfoot RR, Martin DL. 2002.
Dental Research 47:839. The bioarchaeology and taphonomy of
Kosiba SB, Tykot RH, Carlsson D. 2007. Stable violence at Castle Rock and Sand Canyon
isotopes as indicators of change in the food pueblos, southwestern Colorado. American
procurement and food preference of Viking Antiquity 67:486–513.
Age and Early Christian populations on Kuhnke L. 1993. Disease ecologies of the Middle
Gotland (Sweden). Journal of Anthropological East and North Africa. In: Kiple KF, editor. The
Archaeology 26:394–411. Cambridge World History of Human Disease.
Kozłowski T. 2012. Biological State and Life Cambridge, UK: Cambridge University Press;
Conditions of the Population Living in Culmine, p. 453–462.
Pomerania Vistula (10th–13th Century): An Kumar V, Abbas AK, Fausto N, Aster J, editors.
Anthropological Study. Mons Sancti Laurentii, 2010. Robbins and Cotran Pathologic Basis of
Volume 7. Torun, Poland: Nicolaus Copernicus Disease, Tenth Edition. Philadelphia, PA:
University. Saunders Elsevier.
Kreiner A. 1995. Women on the run. American Kupczik K, Dobson CA, Crompton RH, et al. 2009.
Fitness 13:30–33. Masticatory loading and bone adaptation in

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 507

the supraorbital torus of developing macaques. Lallo J, Armelagos GJ, Rose JC. 1978.
American Journal of Physical Anthropology Paleoepidemiology of infectious disease in the
139:193–203. Dickson Mounds population. Medical College
Kyle JH. 1986. Effect of post-burial contamination of Virginia Quarterly 14:17–23.
on the concentrations of major and minor Lallo JW, Rose JC. 1979. Patterns of stress, disease
elements in human bones and teeth – the and mortality in two prehistoric populations
implications for paleodietary research. Journal from North America. Journal of Human
of Archaeological Science 13:403–416. Evolution 8:323–335.
Lacey JM, Anderson JJB, Fujita T, et al. 1991. Lalueza Fox C, González Martin A, Vives Civit S.
Correlates of cortical bone mass among 1996. Cranial variation in the Iberian peninsula
premenopausal and postmenopausal Japanese and the Balearic Islands: inferences about the
women. Journal of Bone and Mineral Research history of the population. American Journal of
6:651–659. Physical Anthropology 99:413–428.
Lahr MM. 1995. Patterns of modern human Lamb SM. 1958. Linguistic prehistory in the Great
diversification: implications for Amerindian Basin. International Journal of American
origins. Yearbook of Physical Anthropology Linguistics 24:95–100.
38:163–198. Lambert JB, Simpson SV, Buikstra JE, Hanson DB.
Lahr MM, Bowman JE. 1992. Palaeopathology of 1983. Electron microprobe analysis of
the Kechipawan site: health and disease in a elemental distribution in excavated human
south-western Pueblo. Journal of femurs. American Journal of Physical
Archaeological Science 19:639–654. Anthropology 62:409–424.
Lahren CH, Berryman HE. 1984. Fracture patterns Lambert JB, Simpson SV, Szpunar CB, Buikstra JB.
and status at Chucalissa (4SY1): a biocultural 1984. Copper and barium as dietary
approach. Tennessee Anthropologist 9:15–21. determinants: the effects of diagenesis.
Lai P, Lovell NC. 1992. Skeletal markers of Archaeometry 26:131–138.
occupational stress in the fur trade: a case Lambert JB, Szpunar CB, Buikstra JE. 1979.
study from a Hudson’s Bay Company fur trade Chemical analysis of excavated human bone
post. International Journal of Osteoarchaeology from Middle and Late Woodland sites.
2:221–234. Archaeometry 21:403–416.
Laine MA. 2002. Effect of pregnancy on Lambert JB, Weydert JM, Williams SR, Buikstra
periodontal and dental health. Acta JE. 1990. Comparison of methods for the
Odontologica Scandinavia 60:267–274. removal of diagenetic material in buried bone.
Laine MA, Tenovuo J, Lehtonen OP, et al. 1988. Journal of Archaeological Science 17:453–468.
Pregnancy-related changes in human whole Lambert PL, Walker PL. 1991. Physical
saliva. Archives of Oral Biology 33:913–917. anthropological evidence for the evolution of
Lalich L, Aufderheide AC. 1991. Lead exposure. In: social complexity in coastal southern
Pfeiffer S, Williamson RF, editors. Snake Hill: California. Antiquity 65:963–973.
An Investigation of a Military Cemetery from Lambert PM. 1993. Health in prehistoric
the War of 1812. Toronto, ON: Dundurn Press; populations of the Santa Barbara Channel
p. 256–262. Islands. American Antiquity 58:509–521.
Lallo JW. 1973. The Skeletal Biology of Three Lambert PM. 1994. War and Peace on the Western
Prehistoric American Indian Societies from Frontier: A Study of Violent Conflict and its
Dickson Mounds. PhD Dissertation, University Correlates in Prehistoric Hunter-Gatherer
of Massachusetts, Amherst, MA. Societies of Coastal Southern California. PhD
Lallo JW, Armelagos GJ, Mensforth RP. 1977. The Dissertation, University of California, Santa
role of diet, disease, and physiology in the Barbara, CA.
origin of porotic hyperostosis. Human Biology Lambert PM. 1997. Patterns of violence in
49:471–483. prehistoric hunter-gatherer societies of coastal

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
508 References

southern California. In: Martin DL, Frayer D, southwestern Colorado: the osteological
editors. Troubled Times: Violence and Warfare evidence from Sleeping Ute Mountain. In:
in the Past. Langhorne, PA: Gordon and Knüsel C, Smith MJ, editors. The Routledge
Breach; p. 77–109. Handbook of the Bioarchaeology of Human
Lambert PM, editor. 2000a. Bioarchaeological Conflict. Oxford, UK: Routledge;
Studies of Life in the Age of Agriculture: p. 308–332.
A View from the Southeast. Tuscaloosa, AL: Lambert PM, Gagnon CM, Billman BR, et al. 2012.
University of Alabama Press. Bone chemistry at Cerro Oreja: a stable isotope
Lambert PM. 2000b. Life on the periphery: health perspective on the development of a regional
in farming communities of interior North economy in the Moche Valley, Peru during the
Carolina and Virginia. In: Lambert PM, editor. Early Intermediate period. Latin American
Bioarchaeological Studies of Life in the Age of Antiquity 23:144–166.
Agriculture: A View from the Southeast. Lambert PM, Leonard BL, Billman BM, et al. 2000.
Tuscaloosa, AL: University of Alabama Press; Response to critique of the claim of
p. 169–194. cannibalism at Cowboy Wash. American
Lambert PM. 2002a. Bioarchaeology at Coweeta Antiquity 65:397–406.
Creek: continuity and change in native health Lane A. 1893. Case of spondylolisthesis associated
and lifeways in protohistoric western North with progressive paraplegia: laminectomy. The
Carolina. Southeastern Archaeology 21:36–48. Lancet 1:991–992.
Lambert PM. 2002b. Rib lesions in a prehistoric Lane DW, Peach DF. 1997. Some observations on
Puebloan sample from southwestern Colorado. the trace element concentrations in human
American Journal of Physical Anthropology dental enamel. Biological Trace Element
117:281–292. Research 60:1–11.
Lambert PM. 2002c. The archaeology of war: a Lane RA, Sublett AJ. 1972. Osteology of social
North American perspective. Journal of organization: residence pattern. American
Archaeological Research 10:207–241. Antiquity 36:186–201.
Lambert PM. 2006. Infectious disease among Lanphear KM. 1990. Frequency and distribution of
enslaved African Americans at Eaton’s enamel hypoplasias in a historic skeletal
Estate, Warren County, North Carolina, sample. American Journal of Physical
ca. 1830–1850. Memórias do Instituto Oswaldo Anthropology 81:35–43.
Cruz 101(Supplement II):107–117. Lanyon LE. 1992. Control of bone architecture by
Lambert PM. 2007. The osteological evidence for functional load bearing. Journal of Bone and
indigenous warfare in North America. In: Mineral Research 7:S369–S375.
Chacon RJ, Mendoza RG, editors. Warfare and Lanyon LE, Baggott DG. 1976. Mechanical
Violence among the Indigenous Peoples of function as an influence on the structure and
North America: Problems in Paradise. Tucson, form of bone. Journal of Bone and Joint
AZ: University of Arizona Press; p. 202–221. Surgery 58-B:436–443.
Lambert PM. 2009. Health versus fitness: Lanyon LE, Bourne S. 1979. The influence of
competing themes in the origins and spread of mechanical function on the development of
agriculture? Current Anthropology remodeling of the tibia. Journal of Bone and
50:603–608. Joint Surgery 61-A:263–273.
Lambert PM. 2012. Ethics and issues in the use of Lanyon LE, Goodship AE, Pye CJ, MacFie JH.
human skeletal remains in paleopathology. In: 1982. Mechanical adaptive bone remodeling.
Grauer AL, editor. A Companion to Journal of Biomechanics 15:142–154.
Paleopathology. Chichester, UK: Wiley- Lanyon LE, Rubin CT. 1984. Static versus
Blackwell; p. 17–33. dynamic loads as an influence on bone
Lambert PM. 2014. Violent injury and death in a remodeling. Journal of Biomechanics
prehistoric farming community of 17:897–905.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 509

Lanyon L, Skerry T. 2001. Postmenopausal Larsen CS. 1998. Gender, health, and activity in
osteoporosis as a failure of bone’s adaptation foragers and farmers in the American
to functional loading: a hypothesis. Journal of Southeast: implications for social organization
Bone and Mineral Research 16:1937–1947. in the Georgia Bight. In: Grauer AL, Stuart-
Larsen CS. 1981. Functional implications of Macadam P, editors. Sex and Gender in
postcranial size reduction on the prehistoric Paleopathological Perspective. Cambridge, UK:
Georgia coast, U.S.A. Journal of Human Cambridge University Press; p. 165–187.
Evolution 10:489–502. Larsen CS, editor. 2001. Bioarchaeology of Spanish
Larsen CS. 1982. The Anthropology of Florida: the Impact of Colonialism. Gainesville,
St. Catherines Island: 3. Prehistoric Human FL: University Press of Florida.
Biological Adaptation. Anthropological Papers Larsen CS. 2006. The agricultural revolution as
of the American Museum of Natural History environmental catastrophe: implications for
57; p. 157–276. health and lifestyle in the Holocene.
Larsen CS. 1983. Deciduous tooth size and Quaternary International 150:12–20.
subsistence change in prehistoric Georgia coast Larsen CS. 2010. Description, hypothesis testing,
populations. Current Anthropology and conceptual advances in physical
24:225–226. anthropology: have we moved on? In: Little
Larsen CS. 1984. Health and disease in prehistoric MA, Kennedy KAR, editors. Histories of
Georgia: the transition to agriculture. In: American Physical Anthropology in the
Cohen MN, Armelagos GJ, editors. Twentieth Century. Lanham, MD: Rowman &
Paleopathology at the Origins of Agriculture. Littlefield; p. 233–241.
Orlando, FL: Academic Press; p. 367–392. Larsen CS. 2012. History of paleopathology in the
Larsen CS. 1985. Dental modifications and tool American Southeast: from pox to population.
use in the western Great Basin. American In Buikstra JE, Roberts CA, Schreiner SM,
Journal of Physical Anthropology 67:393–402. editors: Global History of Paleopathology:
Larsen CS. 1987. Bioarchaeological interpretations Pioneers and Prospects. New York, NY: Oxford
of subsistence economy and behavior from University Press; p. 285–304.
human skeletal remains. Advances in Larsen CS. 2015. Dietary transition in late
Archaeological Method and Theory 10:339–445. Holocene eastern North America: the orofacial
Larsen CS. 1990a. Biological interpretation and record of masticatory function, nutritional
the context for contact. In: Larsen CS, editor. quality, and health in maize farmers. In: Lee-
The Archaeology of Mission Santa Catalina de Thorp J, Katzenberg MA, editors. Oxford
Guale: 2. Biocultural Interpretations of a Handbook of the Archaeology of Diet. New
Population in Transition. Anthropological York, NY: Oxford University Press, in press.
Papers of the American Museum of Natural Larsen CS, Craig J, Sering LE, et al. 1995. Cross
History, No. 68; p. 11–25. Homestead: life and death on the Midwestern
Larsen CS, editor. 1990b. The Archaeology of frontier. In: Grauer AL, editor. Bodies of
Mission Santa Catalina de Guale: 2. Evidence: Reconstructing History through
Biocultural Interpretations of a Population in Skeletal Analysis. New York, NY: Wiley-Liss;
Transition. Anthropological Papers of the p. 139–159.
American Museum of Natural History, No. 68; Larsen CS, Crosby AW, Griffin MC, et al. 2002.
p. 11–150. A biohistory of health and behavior in the
Larsen CS. 1994. In the wake of Columbus: native Georgia Bight: the agricultural transition and
population biology in the postcontact Americas. the impact of European contact. In: Steckel RH,
Yearbook of Physical Anthropology 37:109–154. Rose JC, editors. The Backbone of History:
Larsen CS. 1995. Biological changes in human Health and Nutrition in the Western
populations with agriculture. Annual Review of Hemisphere. New York, NY: Cambridge
Anthropology 24:185–213. University Press; p. 406–439.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
510 References

Larsen CS, Griffin MC, Hutchinson DL, et al. 2001. Larsen CS, Huynh HP, McEwan BG. 1996. Death
Frontiers of contact: bioarchaeology of by gunshot: biocultural implications of trauma
Spanish Florida. Journal of World Prehistory at Mission San Luis. International Journal of
15:69–123. Osteoarchaeology 6:42–50.
Larsen CS, Harn DE. 1994. Health in transition: Larsen CS, Kelly RL, editors. 1995. Bioarchaeology
disease and nutrition in the Georgia Bight. In: of the Stillwater Marsh: Prehistoric Human
Sobolik KD, editor. Paleonutrition: The Diet Adaptation in the Western Great Basin.
and Health of Prehistoric Americans. Southern Anthropological Papers of the American
Illinois University at Carbondale, Center for Museum of Natural History, No. 77; p. 7–170.
Archaeological Investigations, Occasional Larsen CS, Kelly RL, Ruff CB, et al. 2008. Living on
Paper, No. 22; p. 222–234. the margins: biobehavioral adaptations in the
Larsen CS, Hillson SW, Ruff CB, Sadvari JW, western Great Basin. In: Reitz E, Scarry CM,
Garofalo EM. 2013. The human remains II: Scudder SJ, editors. Case Studies in
Interpreting lifestyle and activity in Neolithic Environmental Archaeology. New York, NY:
Çatalhöyük. In: Hodder I, editor. Humans and Springer; p. 161–189.
Landscapes of Çatalhöyük. Los Angeles, CA: Larsen CS, Milner GR, editors. 1994. In the Wake
Cotsen Institute of Archaeology Press; of Contact: Biological Responses to Conquest.
p. 397–412. New York, NY: Wiley-Liss.
Larsen CS, Hutchinson DL. 1992. Dental evidence Larsen CS, Ruff CB. 1991. Biomechanical
for physiological disruption: biocultural adaptation and behavior on the prehistoric
interpretations from the eastern Spanish Georgia coast. In: Powell ML, Bridges PS, Mires
borderlands, U.S.A. In: Goodman A, Capasso AMW, editors. What Mean These Bones?
LL, editors. Recent Contributions to the Study of Studies in Southeastern Bioarchaeology.
Enamel Developmental Defects. Journal of Tuscaloosa, AL: University of Alabama Press;
Paleopathology Monographic Publications, p. 102–113.
No. 2; p. 151–169. Larsen CS, Ruff CB. 1994. The stresses of conquest
Larsen CS, Hutchinson DL. 2010. Osteopathology in Spanish Florida: structural adaptation and
of Carson Desert foragers: reconstructing change before and after contact. In: Larsen CS,
prehistoric lifeways in the western Great Basin. Milner GR, editors. In the Wake of Contact:
In: Hemphill BE, Larsen CS, editor. Biological Responses to Conquest. New York,
Understanding Prehistoric Lifeways in the NY: Wiley-Liss; p. 21–34.
Great Basin Wetlands: Bioarchaeological Larsen CS, Ruff CB. 2011. “An external agency of
Reconstruction and Interpretation. Salt Lake considerable importance”: the stresses of
City, UT: University of Utah Press; p. 184–202. agriculture in the foraging-to-farming
Larsen CS, Hutchinson DL, Schoeninger MJ, Norr transition in eastern North America. In: Pinhasi
L. 2001. Food and stable isotopes in La Florida: R, Stock J, editors. Human Bioarchaeology of
diet and nutrition before and after contact. In: the Transition to Agriculture. Chichester, UK:
Larsen CS, editors. Bioarchaeology of Spanish Wiley-Blackwell; p. 293–315.
Florida: the Impact of Colonialism. Gainesville, Larsen CS, Ruff CB, Griffin MC. 1996. Implications
FL: University Press of Florida; p. 52–81. of changing biomechanical and nutritional
Larsen CS, Hutchinson DL, Stojanowski CM, et al. environments for activity and lifeway in the
2007. Health and lifestyle in Georgia and eastern Spanish borderlands. In: Baker BJ,
Florida: agricultural origins and intensification Kealhofer LL, editors. Bioarchaeology of Native
in regional perspective. In: Cohen MN, Crane- Americans in the Spanish Borderlands.
Kramer GMM, editors. Ancient Health: Skeletal Gainesville, FL: University Press of Florida;
Indicators of Agricultural and Economic p. 95–125.
Intensification. Gainesville, FL: University Larsen CS, Ruff CB, Kelly RL. 1995. Structural
Press of Florida; p. 20–34. analysis of the Stillwater postcranial human

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 511

remains: behavioral implications of articular Papers of the American Museum of Natural


joint pathology and long bone diaphyseal History, 57 (part 4); p. 271–342.
morphology. In: Larsen CS, Kelly RL, editors. Larsen CS, Walker PL. 2005. The ethics of
Bioarchaeology of the Stillwater Marsh: bioarchaeology. In: Turner T, editor. Biological
Prehistoric Human Adaptation in the Western Anthropology and Ethics: From Repatriation to
Great Basin. Anthropological Papers of the Genetic Identity. Albany, NY: State University
American Museum of Natural History, No. 77; of New York Press; p. 111–120.
p. 107–133. Larsen CS, Walker PL. 2010. Bioarchaeology:
Larsen CS, Ruff CB, Schoeninger MJ, Hutchinson health, lifestyle, and society in recent human
DL. 1992. Population decline and extinction in evolution. In: Larsen CS, editor. A Companion
La Florida. In: Verano JW, Ubelaker DH, to Biological Anthropology. Malden, MA:
editors. Disease and Demography in the Wiley-Blackwell; p. 379–394.
Americas. Washington, DC: Smithsonian Larsen CS, Williams LL. 2012. Internationalizing
Institution Press; p. 25–39. physical anthropology: a view of the study of
Larsen CS, Schoeninger MJ, van der Merwe NJ, living human variation from the pages of the
Moore KM, Lee-Thorp JA. 1992. Carbon and American Journal of Physical Anthropology.
nitrogen stable isotopic signatures of human Current Anthropology 53:S139–S151.
dietary change in the Georgia Bight. American Larson LH Jr. 1972. Functional considerations of
Journal of Physical Anthropology 89:197–214. warfare in the Southeast during the Mississippi
Larsen CS, Sering LE. 2000. Inferring iron period. American Antiquity 37:383–392.
deficiency anemia from human skeletal Lasker GW. 1953. The age factor in bodily
remains: the case of the Georgia Bight. In: measurements of adult male and female
Lambert PM, editors. Bioarchaeological Studies Mexicans. Human Biology 25:50–63.
in Life in the Age of Agriculture. Tuscaloosa, Lau EM, Cooper C, Wickham C, Donnan S, Barker
AL: University of Alabama Press; p. 116–133. DJ. 1990. Hip fracture in Hong Kong and
Larsen CS, Shavit R, Griffin MC. 1991. Dental Britain. International Journal of Epidemiology
caries evidence for dietary change: an 19:1119–1121.
archaeological context. In: Kelley MA, Larsen Laughlin WS, Laughlin SB, Beman SB. 1991. Aleut
CS, editors. Advances in Dental Anthropology. kayak-hunter’s hypertrophic humerus. Current
New York, NY: Wiley-Liss; p. 179–202. Research in the Pleistocene 8:55–56.
Larsen CS, Teaford MF, Sandford MK. 1998. Teeth Laughlin WS, Laughlin SB, Beman SB. 1992.
as tools at Tutu: extramasticatory behavior in Kayaker’s hypertrophic humerus. American
prehistoric St. Thomas, U.S. Virgin Islands. In: Journal of Physical Anthropology Supplement
Lukacs JA, editors. Human Dental 14:106.
Development, Morphology, and Pathology: Lavelle CLB. 1968. Anglo-Saxon and modern
A Tribute to Albert A. Dahlberg. University of British teeth. Journal of Dental Research
Oregon Anthropological Papers, No. 54; 47:811–815.
p. 401–420. Lavelle CLB. 1972. A comparison between the
Larsen CS, Teaford MF, Sandford MK. 2002. The mandibles of Romano-British and nineteenth
Tutu teeth: assessing prehistoric health and century periods. American Journal of Physical
lifeway from St. Thomas. In: Righter E, editor. Anthropology 36:213–219.
The Tutu Archaeological Village Site: Lavelle CLB, Moore WJ. 1969. Alveolar bone
A Multidisciplinary Case Study in Human resorption in Anglo-Saxon and seventeenth
Adaptation. New York, NY: Routledge; century mandibles. Journal of
p. 230–249. Periodontological Research 4:70–73.
Larsen CS, Thomas DH. 1982. The Anthropology of Lawrence JS. 1955. Rheumatism in coal miners.
St. Catherines Island. 4. The St. Catherines Part III: Occupational factors. British Journal of
Period Mortuary Complex. Anthropological Industrial Medicine 12:249–261.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
512 References

Lawrence JS. 1961. Rheumatism in cotton bone collagen and bone apatite, and their
operatives. British Journal of Industrial relationship to diet. Journal of Archaeological
Medicine 18:270–276. Science 16:585–599.
Lawrence JS. 1977. Rheumatism in Populations. Le Huray JD, Schutkowski H. 2005. Diet and status
London, UK: William Heinemann Medical during the La Tène period in Bohemia: carbon
Books Ltd. and nitrogen stable isotope analysis of bone
Lawson J. 1967. A New Voyage to Carolina. Chapel collagen from Kutna Hora-Karlov and
Hill, NC: University of North Carolina Press. Radovesice. Journal of Anthropological
Layrisse M, Martinez-Torres C, Roche M. 1968. Archaeology 24:135–147.
Effect of interaction of various foods on iron Leigh RW. 1925. Dental pathology of the Eskimo.
absorption. American Journal of Clinical Dental Cosmos 67:884–898.
Nutrition 21:1175–1183. Leigh RW. 1929. Dental Morphology and
Lazenby R. 2002. Circumferential variation in Pathology of Prehistoric Guam. Memoirs of the
human second metacarpal cortical thickness: Bernice P. Bishop Museum, 11(3).
sex, age, and mechanical factors. Anatomical Leigh RW. 1934. Notes on the Somatology and
Record 267:154–158. Pathology of Ancient Egypt. University of
Lazenby RA, Pfeiffer SK. 1993. Effects of a California Publications in American
nineteenth century below-knee amputation Archaeology and Ethnology, 34(1).
and prosthesis on femoral morphology. Leitman T, Porco T, Blower S. 1993. Leprosy and
International Journal of Osteoarchaeology tuberculosis: the epidemiological consequences
3:19–28. of cross-immunity. International Journal of
Leach S, Lewis M, Chenery C, Muldner G, Eckardt Leprosy and Other Mycobacterial Diseases
H. 2009. Migration and diversity in Roman 61:199–204.
Britain: a multidisciplinary approach to the Leonard WR, DeWalt KM, Stansbury JP, McCaston
identification of immigrants in Roman York, MK. 2000. Influence of dietary quality on the
England. American Journal of Physical growth of highland and coastal Ecuadorian
Anthropology 140:546–561. children. American Journal of Human Biology
LeBlanc SA, Turner CG, Morgan ME. 2008. 12:825–837.
Genetic relationships based on discrete dental Leonard WR, Spencer GJ, Galloway VA, Osipova
traits: Basketmaker II and Mimbres. L. 2002. Declining growth status of indigenous
International Journal of Osteoarchaeology Siberian children in post-Soviet Russia. Human
18:109–130. Biology 74:197–209.
Ledger M, Holtzhausen L-M, Constant D, Morris Lertrit P. Poolsuwan S, Thosarat R, et al. 2008.
AG. 2000. Biomechanical beam analysis of long Genetic history of Southeast Asian populations
bones from a late 18th century slave cemetery in as revealed by ancient and modern human
Cape Town, South Africa. American Journal of mitochondrial DNA analysis. American Journal
Physical Anthropology 112:207–216. of Physical Anthropology 137:425–440.
Lee RB. 1968. What hunters do for a living; or, Lesley BP. 1995. Early evidence of scalping and
how to make out on scarce resources. In: Lee warfare in Late Woodland Illinois. American
RB, DeVore I, editors. Man the Hunter. Chicago, Journal of Physical Anthropology Supplement
IL: Aldine; p. 30–48. 20:134.
Lee RB. 1979. The !Kung San: Men, Women, and Leslie PW, Campbell KL, Campbell BC, Kigondu
Work in a Foraging Society. Cambridge, UK: CS, Kirumbi LW. 1999. Fecundity and fertility.
Cambridge University Press. In: Little A, Leslie PW, editors. Turkana Herders
Lee-Thorp JA. 2008. On isotopes and old bones. of the Dry Savanna: Ecology and
Archaeometry 6:925–950. Biobehavioural Response of Nomads to an
Lee-Thorp JA, Sealy JC, van der Merwe NJ. 1989. Uncertain Environment. New York, NY: Oxford
Stable carbon isotope ratio differences between University Press; p. 249–278

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 513

Lessa A. 2011. Spondylolysis and lifestyle among Li X, Kolltveit KM, Tronstad L, Olsen I. 2000.
prehistoric coastal groups from Brazil. Systemic diseases caused by oral infection.
International Journal of Osteoarchaeology Clinical Microbiology Reviews 13:547–558.
21:660–668. Li Y, Navia JM, Bian J-Y. 1995. Prevalence and
Lester CW, Shapiro HL. 1968. Vertebral arch distribution of developmental enamel defects
defects in the lumbar vertebrae of pre-historic in primary dentition of Chinese children
American Eskimos: a study of skeletons in the 3–5 years old. Community Dentistry and Oral
American Museum of Natural History, chiefly Epidemiology 23:72–79.
from Point Hope, Alaska. American Journal of Lichtheim M. 1973. Ancient Egyptian Literature,
Physical Anthropology 28:43–48. Volume 1: The Old and Middle Kingdoms.
Leverett DH. 1982. Fluorides and the changing Berkeley, CA: University of California Press.
prevalence of dental caries. Science 217:26–30. Lidén K. 1995. Megaliths, agriculture, and social
Levy LF. 1968. Porter’s neck. British Medical complexity: a diet study of two Swedish
Journal 2:16–19. Megalith populations. Journal of
Lewis B. 1994. Treponematosis and Lyme Anthropological Archaeology 14:404–417.
borreliosis connections: explanation for Lieberman DE. 1995. Testing hypotheses about
Tchefuncte disease syndromes? American recent human evolution from skulls. Current
Journal of Physical Anthropology Anthropology 36:159–197.
93:455–475. Lieberman DE. 2011. The Evolution of the Human
Lewis JE. 2008. Identifying sword marks on bone: Head. Cambridge, MA: Harvard University
criteria for distinguishing between cut marks Press.
made by different classes of bladed weapons. Lieberman DE, Krovitz GE, Yates FW, Devlin M,
Journal of Archaeological Science St. Claire M. 2004. Effects of food processing
35:2001–2008. on masticatory strain and craniofacial growth
Lewis ME. 2002. Impact of industrialization: in a retrognathic face. Journal of Human
comparative study of child health in four sites Evolution 46:655–677.
from medieval and postmedieval England Lieberman DE, Polk JD, Demes B. 2004. Predicting
(A.D. 850–1859). American Journal of Physical long bone loading from cross-sectional
Anthropology 119:211–223. geometry. American Journal of Physical
Lewis ME. 2007. The Bioarchaeology of Children: Anthropology 123:156–171.
Perspectives from Biological and Forensic Liesau von Lettow-Vorbeck C, Pastor Abascal I.
Anthropology. Cambridge, UK: Cambridge 2003. The Ciempozuelos necropolis skull: a
University Press. case of double trepanation? International
Lewis ME. 2014. Sticks and stones: exploring the Journal of Osteoarchaeology 13:213–221.
nature and significance of child trauma in the Lieverse AR, Bazaliiskii VI, Goriunova OI, Weber
past. In: Knüsel C, Smith MJ, editors. The AW. 2009. Upper limb musculoskeletal stress
Routledge Handbook of the Bioarchaeology of markers among middle Holocene foragers of
Human Conflict. London, UK: Routledge; Siberia’s Cis-Baikal region. American Journal
p. 39–64. of Physical Anthropology 158:458–472.
Lewis ME, Roberts CA, Manchester K. 1995. Lieverse AR, Link DW, Bazaliiskiy VI, Goriunova
Comparative study of the prevalence of OI, Weber AW. 2007. Dental health indicators
maxillary sinusitis in later Medieval urban and of hunter-gatherer adaptation and cultural
rural populations in northern England. change in Siberia’s Cis-Baikal. American
American Journal of Physical Anthropology Journal of Physical Anthropology
98:497–506. 134:323–339.
Lewis TMN, Lewis MK. 1961. Eva: An Archaic Lieverse AR, Stock JT, Katzenberg MA, Haverkort
Site. Knoxville, TN: University of Tennessee CM. 2011. The bioarchaeology of habitual
Press. activity and dietary change in the Siberian

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
514 References

middle Holocene. In: Pinhasi R, Stock JT, Agriculture. Chichester, UK: Wiley-Blackwell;
editors. Human Bioarchaeology of the p. 385–402.
Transition to Agriculture. Chichester, UK: John Linderholm A, Kjellstrom A. 2011. Stable isotope
Wiley & Sons; p. 265–291. analysis of a medieval skeletal sample
Lieverse AR, Weber AW, Bazaliiskiy VI, Goriunova indicative of systemic disease from Sigtuna,
OI, Savel’ev NA. 2007. Osteoarthritis in Sweden. Journal of Archaeological Science
Siberia’s Cis-Baikal: skeletal indicators of 38:925–933.
hunter-gatherer adaptation and cultural Linkosalo E, Markkanen H. 1985. Dental erosions
change. American Journal of Physical in relation to lactovegetarian diet.
Anthropology 132:1–16. Scandinavian Journal of Dental Research
Lightfoot E, Boneva B, Miracle PT, Šlaus M, 93:436–441.
O’Connell TC. 2011. Exploring the Mesolithic Liston MA, Baker BJ. 1996. Reconstructing the
and Neolithic transition in Croatia through massacre at Fort William Henry, New York.
isotopic investigations. Antiquity International Journal of Osteoarchaeology
85:73–86. 6:28–41.
Lightfoot E, Šlaus M, O’Connell TC. 2012. Little EA, Schoeninger MJ. 1995. The Late
Changing cultures, changing cuisines: cultural Woodland diet on Nantucket Island and
transitions and dietary change in Iron Age, the problem of maize in coastal New
Roman, and Early Medieval Croatia. American England. American Antiquity 60:351–368.
Journal of Physical Anthropology 148:543–556. Little MA, Sussman RW. 2010. History of
Lightfoot KG, Martinez A. 1995. Frontiers and biological anthropology. In: Larsen CS, editor.
boundaries in archaeological perspective. A Companion to Biological Anthropology.
Annual Review of Anthropology 24:471–492. Malden, MA: Wiley-Blackwell; p. 13–38.
Likins RC, McCann HG, Posner AS, Scott DB. Littleton J. 2005. Invisible impacts but long-term
1960. Comparative fixation of calcium and consequences: hypoplasia and contact in
strontium by synthetic hydroxyapatite. Journal central Australia. American Journal of Physical
of Biological Chemistry 235:2152–2156. Anthropology 126:295–304.
Lillie MC. 1996. Mesolithic and Neolithic Littleton J. 2011. Moving from the canary in the
populations of Ukraine: indications of diet coalmine. In: Agarwal SC, Glencross BA,
from dental pathology. Current Anthropology editors. Social Bioarchaeology. Chichester, UK:
37:135–142. Wiley-Blackwell; p. 361–389.
Lillie M, Budd C. 2011. The Mesolithic–Neolithic Littleton J, Frohlich B. 1993. Fish-eaters and
transition in eastern Europe: integrating stable farmers: dental pathology in the Arabian Gulf.
isotope studies of diet with palaeopathology to American Journal of Physical Anthropology
identify subsistence strategies and economy. 92:427–447.
In: Pinhasi R, Stock JT, editors. Human Littleton J, Scott R, McFarlane G, Walshe K. 2013.
Bioarchaeology of the Transition to Agriculture. Hunter-gatherer variability: dental wear in
Chichester, UK: Wiley-Blackwell; p. 43–62. South Australia. American Journal of Physical
Lillie M, Budd C, Potekhina I. 2011. Stable isotope Anthropology 152:273–286.
analysis of prehistoric populations from the Liu W, Zhang QC, Wu XJ, Zhu H. 2010. Tooth
cemeteries of the Middle and Lower Dnieper wear and dental pathology of the Bronze–Iron
Basin, Ukraine. Journal of Archaeological Age people in Xinjiang, northwest China:
Science 38:57–68. implications for their diet and lifestyle. Homo
Linderholm A. 2011. The genetics of the Neolithic 61:102–116.
transition: new light on differences between Lobdell JE. 1984. Harris lines: markers of
hunter-gatherers and farmers in southern nutrition and disease at prehistoric
Sweden. In: Pinhasi R, Stock JT, editors. Utqiagvik Village. Arctic Anthropology
Human Bioarchaeology of the Transition to 21:109–116.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 515

Lobdell JE. 1988. Harris lines: markers of southeast Anatolia. American Journal of
nutritional stress in late prehistoric and contact Physical Anthropology 131:181–193.
period Eskimo postcranial remains. In: Shaw Lovász G, Schultz M, Gödde J, et al. 2014. Skeletal
RD, Harritt RK, Dumond DE, editors. The Late manifestations of infantile scurvy in a late
Prehistoric Development of Alaska’s Native medieval anthropological series from Hungary.
People. Aurora IV. Anchorage, AK: Alaska Anthropological Science 121:173–185.
Anthropological Association; p. 47–55. Love JC, Derrick SM, Wiersema JM. 2011. Skeletal
Loe H. 1965. Periodontal changes in pregnancy. Atlas of Child Abuse. New York, NY: Springer.
Journal of Periodontology 36:209–217. Lovejoy CO, Burstein AH, Heiple KG. 1976. The
Loe H, Silness J. 1963. Periodontal disease in biomechanical analysis of bone strength: a
pregnancy: I. Prevalence and severity. Acta method and its application to platycnemia.
Odontologica Scandinavica 21:533–551. American Journal of Physical Anthropology
Loesche WJ, Grossman NS. 2001. Periodontal 44:489–506.
disease as a specific, albeit chronic, infection: Lovejoy CO, Heiple KG. 1981. The analysis of
diagnosis and treatment. Clinical Microbiology fractures in skeletal populations, with an
Reviews 14:727–752. example from the Libben site, Ottawa County,
Lombardi AV, Bailit HL. 1972. Malocclusion in the Ohio. American Journal of Physical
Kwaio, a Melanesian group on Malaito, Anthropology 55:529–541.
Solomon Islands. American Journal of Physical Lovejoy CO, Meindl RS, Mensforth RP, Barton, TJ.
Anthropology 36:283–294. 1985. Multifactorial determination of skeletal
Lorentzon M, Mellstrom D, Ohlsson C. 2005. age at death: a method and blind tests of its
Association of amount of physical activity with accuracy. American Journal of Physical
cortical bone size and trabecular volumetric Anthropology 68:1–14.
BMD in young adult men: the GOOD study. Lovejoy CO, Meindl RS, Pryzbeck TR, et al. 1977.
Journal of Bone and Mineral Research Paleodemography of the Libben site, Ottawa
20:1936–1943. County, Ohio. Science 198:291–293.
Lorenz JG, Smith DG. 1996. Distribution of four Lovejoy CO, Mensforth RP, Armelagos GJ. 1982.
founding mtDNA haplogroups among native Five decades of skeletal biology as reflected in
North Americans. American Journal of the American Journal of Physical
Physical Anthropology 101:307–323. Anthropology. In: Spencer F, editor. A History
Loring S, Prokopec M. 1994. A most peculiar man: of American Physical Anthropology:
the life and times of Aleš Hrdlička. In: Bray TL, 1930–1980. New York, NY: Academic Press;
Killion TW, editors. Reckoning With the Dead: p. 329–336.
The Larsen Bay Repatriation and the Lovejoy CO, Russell KF, Harrison ML. 1990. Long
Smithsonian Institution. Washington, DC: bones growth velocity in the Libben
Smithsonian Institution Press; p. 26–40. population. American Journal of Human
Lorkiewicz W. 2011. Nonalimentary tooth use in Biology 2:533–541.
the Neolithic population of the Lengyel Culture Lovejoy CO, Trinkaus E. 1980. Strength and
in central Poland. American Journal of robusticity of the Neanderthal tibia. American
Physical Anthropology 144:538–551. Journal of Physical Anthropology 53:465–470.
Lorkiewicz W, Stolarczyk H, Smiszkiewicz- Lovell NC. 1994. Spinal arthritis and physical
Skwarska A, Zadzinska E. 2005. An interesting stress at Bronze Age Harappa. American
case of prehistoric trepanation from the Franki Journal of Physical Anthropology 93:149–164.
Suchodolskie site. International Journal of Lovell NC. 1997. Trauma analysis in
Osteoarchaeology 15:115–123. paleopathology. Yearbook of Physical
Losch S, Grupe G, Peters J. 2006. Stable isotopes Anthropology 40:139–170.
and dietary adaptations in humans and Lovell NC. 2008. Analysis and interpretation of
animals at Pre-Pottery Neolithic Nevali Cori, skeletal trauma. In: Katzenberg MA, Saunders

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
516 References

SR, editors. Biological Anthropology of the short-term subsistence shift. In: Cohen MN,
Human Skeleton, Second Edition. New York, Crane-Kramer GMM, editors. Ancient Health:
NY: John Wiley & Sons; p. 341–386. Skeletal Indicators of Agricultural and
Lovell NC, Lai P. 1994. Lifestyle and health of Economic Intensification. Gainesville, FL:
voyageurs in the Canadian fur trade. In: University Press of Florida; p. 237–249.
Herring A, Chan L, editors. Strength and Lukacs JR. 2007b. Dental trauma and antemortem
Diversity: A Reader in Physical Anthropology. tooth loss in prehistoric Canary Islanders:
Toronto, ON: Canadian Scholars’ Press; prevalence and contributing factors.
p. 327–343. International Journal of Osteoarchaeology
Lowie RH. 1954. Indians of the Plains. New York, 17:157–173.
NY: McGraw Hill Book Company. Lukacs JR. 2008. Fertility and agriculture
Lu K. 1977. Dental condition of two tribes of accentuate sex differences in dental caries
Taiwan aborigines–Ami and Atayal. Journal of rates. Current Anthropology 49:901–914.
Dental Research 56:117–126. Lukacs JR. 2012. Oral health in past populations:
Lubell D, Jackes M, Schwarcz H, Knyf M, context, concepts and controversies. In: Grauer
Meiklejohn C. 1994. The Mesolithic–Neolithic A, editor. A Companion to Paleopathology.
transition in Portugal: isotopic and dental Chichester, UK: Wiley-Blackwell;
evidence of diet. Journal of Archaeological p. 553–581.
Science 21:201–216. Lukacs JR, Hemphill BE. 1990. Traumatic
Lucas PW. 2004. Dental Functional Morphology: injuries of prehistoric teeth: new evidence
How Teeth Work. Cambridge, UK: Cambridge from Baluchistan and Punjab provinces,
University Press. Pakistan. Anthropologischer Anzeiger
Lukacs JR. 1983. Dental anthropology and the 48:351–363.
origin of two Iron Age populations of northern Lukacs JR, Hemphill BE. 1991. The dental
Pakistan. Homo 34:1–15. anthropology of prehistoric Baluchistan: a
Lukacs JR. 1985. Tooth size variation in morphometric approach to the peopling of
prehistoric India. American Anthropologist south Asia. In: Kelley MA, Larsen CS, editors.
87:1–15. Advances in Dental Anthropology. New York,
Lukacs JR. 1989. Biological affinities from dental NY: Wiley-Liss; p. 77–119.
morphology: the evidence from Neolithic Lukacs JR, Hemphill BE. 1993. Odontometry and
Mehrgarh. In: Kenoyer JM, editor. Old biologic affinities among South Asians: an
Problems and New Perspectives in the analysis of three ethnic groups from northwest
Archaeology of South Asia. Wisconsin India. Human Biology 65:279–325.
Archaeological Reports, No. 2; p. 75–88. Lukacs JR, Joshi MR. 1992. Enamel hypoplasia
Lukacs JR. 1990. On hunter-gatherers and their prevalence in three ethnic groups of northwest
neighbors in prehistoric India: contact and India: a test of daughter neglect and a
pathology. Current Anthropology 31:183–186. framework for the past. In: Goodman AH,
Lukacs JR. 1992. Dental paleopathology and Capasso LL, editors. Recent Contributions to the
agricultural intensification in South Asia: Study of Enamel Developmental Defects.
new evidence from Bronze Age Harappa. Journal of Paleopathology, Monographic
American Journal of Physical Anthropology Publications, No. 2; p. 359–372.
87:133–150. Lukacs J, Minderman L. 1992. Dental pathology
Lukacs JR. 1996. Sex differences in dental and agricultural intensification from Neolithic
caries rates with the origin of agriculture to Chalcolithic periods at Mehrgarh
in South Asia. Current Anthropology (Baluchistan, Pakistan). In: Jarrige C, editor.
37:147–153. South Asian Archaeology 1989.
Lukacs JR. 2007a. Climate, subsistence and health Monographs in World Archaeology, No. 14;
in prehistoric India: the biological impact of a p. 167–179.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 517

Lukacs JR, Pal JN. 1993. Mesolithic subsistence in Luz B, Cormie AB, Schwarcz HP. 1990. Oxygen
north Indian: inferences from dental attributes. isotope variations in phosphate of deer bones.
Current Anthropology 34:745–765. Geochimica et Cosmochimica Acta
Lukacs JR, Pal JN. 2003. Skeletal variation among 54:1723–1728.
Mesolithic people of the Ganga Plains. Asian Lynnerup N. 1998. The Greenland Norse: a
Perspectives 42:329–351. Biological-Anthropological Study. Meddelelser
Lukacs JR, Pastor RF. 1988. Activity-induced Om Grønland, Man & Society, 24.
patterns of dental abrasion in prehistoric Lynnerup N. 2001. Cranial thickness in relation to
Pakistan: evidence from Mehrgarh and age, sex, and general body build in a Danish
Harappa. American Journal of Physical forensic sample. Forensic Science International
Anthropology 76:377–398. 117:45–51.
Lukacs JR, Schultz M, Hemphill BE. 1989. Dental Lynnerup N, Boldsen J. 2012. Leprosy (Hansen’s
pathology and dietary patterns in Iron Age disease). In: Grauer AL, editor. A Companion to
northern Pakistan. In: Sorensen P, Frifelt K, Paleopathology. Chichester, UK: Wiley-
editors. South Asian Archaeology 1985. Blackwell; p. 458–471.
London, UK: Curzon Press; p. 475–496. Lynnerup N, Jacobsen JCB. 2003. Age and fractal
Lukacs JR, Thompson LM. 2008. Dental caries dimensions of human sagittal and coronal
prevalence by sex in prehistory: magnitude sutures. American Journal of Physical
and meaning. In: Irish JD, Nelson GC, editors. Anthropology 121:332–336.
Technique and Application in Dental Lynnerup N, Mikkelsen H, Poulsen K, et al. 2002.
Anthropology. Cambridge, UK: Cambridge A firearms fatality from 14th century Denmark.
University Press; p. 136–177. Journal of Paleopathology 13:11–16.
Lumbreras LG. 1974. The Peoples and Cultures of Lysell L. 1958. A biometrical study of occlusion
Ancient Peru. Washington, DC: Smithsonian and dental arches in a series of Medieval skulls
Institution Press. from northern Sweden. Acta Odontologica
Lundström A. 1948. Tooth Size and Occlusion in Scandinavica 16:177–203.
Twins. Basel, Switzerland: S. Karger. Ma PH, Teaford MF. 2010. Diet reconstruction in
Lundström A. 1963. Tooth morphology as a basis antebellum Baltimore: insights from dental
for distinguishing monozygotic and dizygotic microwear analysis. American Journal of
twins. American Journal of Human Genetics Physical Anthropology 141:571–582.
15:34–43. Maat GJR, Mastwijk RW. 2000. Avulsion injuries
Lundström A, Lysell L. 1953. An anthropological of vertebral endplates. International Journal of
examination of a group of Mediaeval Danish Osteoarchaeology 10:142–152.
skulls. Acta Odontologica Scandinavica Maat GJR, Van der Velde EA. 1987. The caries–
11:111–128. attrition competition. International Journal of
Lundy JK. 1981. Spondylolysis of the lumbar Anthropology 2:281–292.
vertebrae in a group of prehistoric Upper Puget Macchiarelli R, Bondioli L. 1986. Post-Pleistocene
Sound Indians at Birch Bay, Washington. In: reductions in human dental structure: a
Cybulski JS, editor. Contributions to Physical reappraisal in terms of increasing population
Anthropology, 1978–1980. Archaeological density. Human Evolution 1:405–417.
Survey of Canada, Mercury Series Paper, Macdonald HM, Cooper DML, McKay HA. 2009.
No. 106; p. 107–114. Anterior-posterior bending strength at the
Lunt DA. 1969. An odontometric study of tibial shaft increases with physical activity in
Medieval Danes. Acta Odontologica boys: evidence for non-uniform geometric
Scandinavica 27:55–113. adaptation. Osteoporosis International
Lux B. 1994. Cult or conflict? The human bones from 20:61–70.
the area of the sea trade town Ralswiek on the isle Mack ME, Goodman AH, Blakey ML, Mayes A.
of Rugen (8th to 10th c.). Homo 45:31–50. 2009. Odontological indicators of disease, diet,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
518 References

and nutrition inadequacy. In: Blakey ML, Archaeological Reports, International Series,
Rankin H, editors. The Skeletal Biology of the No. 1418.
New York African Burial Ground, Part I. Mahoney P. 2006. Dental microwear from
Washington, DC: Howard University Press; Natufian hunter-gatherers and early Neolithic
p. 157–168. farmers: comparisons within and between
Magennis AL. 1986. The physical anthropology of samples. American Journal of Physical
the Indian Neck ossuary. In: McManamon FP, Anthropology 130:308–319.
Bradley JW, Magennis A, editors. The Indian Mahoney P. 2007. Human dental microwear from
Neck Ossuary: Chapters in the Archeology of Ohalo II (22,500–23,500 cal BP), southern
Cape Cod, V. National Park Service, Division of Levant. American Journal of Physical
Cultural Resources, North Atlantic Regional Anthropology 132:489–500.
Office, Cultural Resources Management Study, Maier AW, Orban B. 1949. Gingivitis in
No. 17; p. 49–183. pregnancy. Oral Surgery, Oral Medicine, and
Maggiano C, Dupras T, Biggerstaff J. 2003. Oral Pathology 2:334–373.
Ancient antibiotics: evidence of tetracycline in Malhi RS, Eshleman JA, Greenberg JA, et al. 2002.
human and animal bone from Kellis. In: Mills The structure and diversity within New
AJ, Hope CA, editors. The Dakhleh Oasis World mitochondrial DNA haplogroups:
Monograph, Volume 3. Oxford, UK: Oxbow implications for the prehistory of North
Books; p. 331–344. America. American Journal of Human Genetics
Maggiano C, Dupras T, Schultz M, Biggerstaff J. 70:905–919.
2006. Spectral and photobleaching analysis Malhi RS, Mortensen HM, Johnson JR, Gorodezky
using confocal laser scanning microscopy: a C, Smith DG. 2003. Native American mtDNA
comparison of modern and archaeological prehistory in the American Southwest.
bone fluorescence. Molecular and Cellular American Journal of Physical Anthropology
Probes 20:154–162. 120:108–124.
Maggiano C, Dupras T, Schultz M, Biggerstaff J. Malina RM, Pena Reyes ME, Little BB. 2010.
2009. Confocal laser scanning microscopy: a Secular change in heights of indigenous adults
flexible tool for simultaneous polarization and from a Zapotec-speaking community in
three-dimensional fluorescence imaging of Oaxaca, southern Mexico. American Journal of
archaeological compact bone. Journal of Physical Anthropology 141:463–475.
Archaeological Science 36:2392–2401. Malina RM, Pena Reyes ME, Tan SK, et al. 2005.
Maggiano IS, Schultz M, Kierdorf H, et al. 2008. Secular change in height, sitting height, and
Cross-sectional analysis of long bones, leg length in rural Oaxaca, southern Mexico.
occupational activities and long-distance trade Annals of Human Biology 31:1968–2000.
of the Classic Maya from Xcambo – Malthus TR. 1798. An Essay on the Principle of
archaeological and osteological evidence. Population, edited by Philip Appleman, 2004.
American Journal of Physical Anthropology New York, NY: WW Norton.
136:470–477. Ma’luf RN, Zein WM, El Dairi MA, Bashshur ZF.
Magilton J, Lee F, Boylston A. 2008. Lepers 2002. Biolateral subperiosteal orbital
outside the Gate: Excavations at the Cemetery hematomas and Henoch–Schonlein purpura.
of the Hospital of St. James and St. Mary Archives of Ophthalmology 120:1398–1399.
Magdalene, Chichester, 1986–1987 and 1993. Malville NJ. 2008. Stature of ancestral Pueblo
Bootham, UK: Council for British Archaeology. populations: population density, social
Magnusun PB. 1942. Fractures. Philadelphia, PA: stratification, and dietary protein. In: Stodder
Lea & Febiger. ALW, editor. Reanalysis and Reinterpretation
Mahoney P. 2005. Dental Microwear in Natufian in Southwestern Bioarchaeology. Arizona State
Hunter-Gatherers and Pre-Pottery Neolithic University Anthropological Research Papers,
Farmers from Northern Israel. British No. 59; p. 105–126.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 519

Manchester K. 1981. A leprous skeleton of the 7th of a Neolithic sample from Italy. American
century from Eccles, Kent, and the present Journal of Physical Anthropology 137:188–200.
evidence of leprosy: evidence of leprosy in Marchi D, Sparacello V, Shaw C. 2011. Mobility
early Britain. Journal of Archaeological Science and lower limb robusticity of a pastoralist
8:205–209. Neolithic population from north-western Italy.
Manchester K. 1982. Spondylolysis and In: Pinhasi R, Stock JT, editors. Human
spondylolisthesis in two Anglo-Saxon Bioarchaeology of the Transition to Agriculture.
skeletons. Paleopathology Newsletter 37:9–12. Chichester, UK: John Wiley & Sons; p. 317–346.
Manchester K. 1983. The Archaeology of Disease. Marcus R, Drinkwater B, Dalsky G, et al. 1992.
Bradford, UK: University of Bradford Press. Osteoporosis and exercise in women.
Manchester K. 1991. Tuberculosis and leprosy: Medicine and Science in Sports and Exercise
evidence for interaction of disease. In: Ortner 24:S301–S307.
DJ, Aufderheide AC, editor. Human Marden K, Ortner DJ. 2011. A case of
Paleopathology: Current Synthesis and Future treponematosis from pre-Columbian Chaco
Options. Washington, DC: Smithsonian Canyon, New Mexico. International Journal of
Institution Press; p. 23–35. Osteoarchaeology 21:19–31.
Manchester K. 1993. Unusual pathological Maresh MM. 1970. Measurements from
condition in the lower extremities of a skeleton roentgenograms. In: McCammon RW, editor.
from ancient Israel. American Journal of Human Growth and Development. Springfield,
Physical Anthropology 91:249–250. IL: Charles C Thomas; p. 157–199.
Manchester K, Elmhirst OEC. 1980. Forensic Margerison BJ, Knüsel CJ. 2002.
aspects of an Anglosaxon injury. Ossa Paleodemographic comparison of a
7:179–188. catastrophic and an attritional death
Manchester K, Roberts CA. 1989. The assemblage. American Journal of Physical
palaeopathology of leprosy in Britain: a review. Anthropology 119:134–143.
World Archaeology 21:265–272. Margetts EL. 1967. Trepanation of the skull by the
Mandell GL, Douglas RG, Bennett JG. 1990. medicine-men of primitive cultures, with
Principles and Practice of Infectious Diseases, particular reference to present-day native east
Third Edition. New York, NY: Churchill African practice. In: Brothwell D, Sandison AT,
Livingstone. editors. Diseases in Antiquity. Springfield, IL:
Manek NJ, Spector TD. 2003. Evidence for Charles C. Thomas; p. 673–701.
inheritance of osteoarthritis. In: Brandt KD, Marion LR. 1996. Dentistry in ancient Egypt.
Doherty M, Lohmander LS, editors. Journal of the History of Dentistry 44:15–17.
Osteoarthritis, Second Edition. New York, NY: Mariotti V, Dutour O, Belcastro MG, Facchini F,
Oxford University Press; p. 25–31. Brasili P. 2005. Probable early presence of
Mann RW, Owsley DW, Reinhard KJ. 1994. Otitis leprosy in Europe in a Celtic skeleton of the
media, mastoiditis, and infracranial lesions in 4th–3rd century BC (Casalvecchio di Reno,
two Plains Indian children. In: Owsley DW, Bologna, Italy). International Journal of
Jantz RL, editors. Skeletal Biology in the Great Osteoarchaeology 15:311–325.
Plains: Migration, Warfare, Health, and Marks J. 1995. Human Biodiversity: Genes, Race,
Subsistence. Washington, DC: Smithsonian and History. New York, NY: Aldine de Gruyter.
Institution Press; p. 131–146. Marks M, Rose JC, Buie EL. 1988. Bioarchaeology
Mansilla J, Pijoan CM. 1995. A case study of of Seminole Sink. In: Seminole Sink:
congenital syphilis during the Colonial period Excavation of a Vertical Shaft Tomb, Val Verde
in Mexico City. American Journal of Physical County, Texas. Plains Anthropologist Memoir,
Anthropology 97:187–195. No. 22; p. 75–118.
Marchi D. 2008. Relationships between lower limb Marlar RA, Leonard BL, Billman BM, Lambert PM,
cross-sectional geometry and mobility: the case Marlar JE. 2000. Biochemical evidence of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
520 References

cannibalism at a prehistoric Puebloan site in Martin DL, Anderson CP, editors. 2014.
southwestern Colorado. Nature 407:74–78. Bioarchaeological and Forensic Perspectives on
Márquez Morfín L, del Ángel A. 1997. Height Violence: How Violent Death is Interpreted
among prehispanic Maya of the Yucatan from Skeletal Remains. Cambridge, UK:
peninsula: a reconsideration. In: Whittington Cambridge University Press.
SL, Reed DM, editors. Bones of the Maya: Martin DL, Armelagos GJ. 1979. Morphometrics of
Studies of Ancient Skeletons. Washington, DC: compact bone: an example from Sudanese
Smithsonian Institution Press; p. 51–61. Nubia. American Journal of Physical
Márquez Morfín L, McCaa R, Storey R, del Ángel Anthropology 51:571–578.
A. 2002. Health and nutrition in pre-Hispanic Martin DL, Armelagos GJ. 1985. Skeletal
Mesoamerica. In: Steckel RH, Rose JC, editors. remodeling and mineralization as indicators of
The Backbone of History: Health and Nutrition health: an example from prehistoric Sudanese
in the Western Hemisphere. New York, NY: Nubia. Journal of Human Evolution
Cambridge University Press; p. 307–338. 14:527–537.
Márquez Morfín L, Storey R. 2007. From early Martin DL, Frayer DW, editors. 1997. Troubled
village to regional center in Mesoamerica: an Times: Violence and Warfare in the Past.
investigation of lifestyles and health. In: Amsterdam, the Netherlands: Gordon and
Cohen MN, Crane-Kramer GMM, editors. Breach Publishers.
Ancient Health: Skeletal Indicators of Martin DL, Goodman AH, Armelagos GJ. 1985.
Agricultural and Economic Intensification. Skeletal pathologies as indicators of quality
Gainesville, FL: University Press of Florida; and quantity of diet. In: Gilbert RI, Mielke JH,
p. 80–91. editors. The Analysis of Prehistoric Diets.
Marshall D, Johnell O, Wedel H. 1996. Meta- Orlando, FL: Academic Press; p. 227–279.
analysis of how well measures of bone mineral Martin DL, Goodman AH, Armelagos GJ,
density predict occurrence of osteoporotic Magennis AL. 1991. Black Mesa Anasazi
fractures. British Medical Journal Health: Reconstructing Life from Patterns of
312:1254–1259. Death and Disease. Southern Illinois University
Marshall J. 1931. Mohenjo-daro and the Indus at Carbondale, Center for Archaeological
Civilization. London, UK: A. Probsthain. Investigations, Occasional Paper, No. 14.
Marshall WA. 1968. Problems in relating the Martin DL, Harrod RP, Pérez VR, editors. 2012. The
presence of transverse lines in the radius to the Bioarchaeology of Violence. Gainesville, FL:
occurrence of disease. In: Brothwell DR, editor. University Press of Florida.
The Skeletal Biology of Earlier Human Martin DL, Harrod RP, Pérez VR, editors. 2013.
Populations. London, UK: Pergamon; Bioarchaeology: An Integrated Approach to
p. 245–261. Working with Human Remains. New York, NY:
Martin DL. 2008. Reanalysis of trauma in the La Springer.
Plata Valley (A.D. 900–1300): strategic social Martin RB. 2007. The importance of mechanical
violence and the bioarchaeology of captivity. loading in bone biology and medicine. Journal
In: Stodder ALW, editor. Reanalysis and of Musculoskeletal and Neuronal Interactions
Reinterpretation in Southwestern 7:48–53.
Bioarchaeology. Arizona State University, Martin RB, Atkinson PJ. 1977. Age and sex-
Anthropological Research Papers, No. 59; related changes in the structure and strength of
p. 166–183. the human femoral shaft. Journal of
Martin DL, Akins NJ, Goodman AH, Toll HW, Biomechanics 10:223–231.
Swedlund AC. 2001. Harmony and Discord: Martin RB, Burr DB. 1984. Non-invasive
Bioarchaeology. Santa Fe: Office of measurement of long bone cross-sectional
Archaeological Studies, Archaeology Notes, moment of inertia by photon absorptiometry.
No. 242. Journal of Biomechanics 17:195–201.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 521

Martin RB, Burr DB, Sharkey NA. 1998. Skeletal of Southeast Asia. Cambridge, UK: Cambridge
Tissue Mechanics. New York, NY: Springer. University Press; p. 33–58.
Martini F, Ober WC. 2001. Fundamentals of Matsumura H. 2011. Quantitative cranio-
Anatomy and Physiology. New York, NY: morphology at Man Bac. In: Oxenham MF,
Prentice-Hall. Matsumura H, Dung NK, editors. Man Bac: The
Martorell R, Yarbrough C, Lechtig A, Habicht J-P, Excavation of a Neolithic Site in Northern
Klein RE. 1977. Diarrheal diseases and growth Vietnam – The Biology. Canberra, Australia:
retardation in preschool Guatemalan children. ANU E Press; p. 21–31.
American Journal of Physical Anthropology Matsumura H, Hudson MJ. 2005. Dental
43:341–346. perspectives on the population history of
Maschner HDG. 1992. The Origins of Hunter and Southeast Asia. American Journal of Physical
Gatherer Sedentism and Political Complexity: Anthropology 127:182–209.
A Case Study from the Northern Northwest Matthews W. 1893. The human bones of the
Coast. PhD Dissertation, University of Hemenway Collection in the United States
California, Santa Barbara, CA. Army Medical Museum at Washington.
Maschner HDG, Reedy-Maschner KL. 1998. Raid, Proceedings of the National Academy of
retreat, defend (repeat): the archaeology and Sciences, 6.
ethnohistory of warfare on the north Pacific May RL, Goodman AH, Meindl RS. 1993. Response
rim. Journal of Anthropological Archaeology of bone and enamel formation to nutritional
17:19–51. supplementation and morbidity among
Massler M, Schour I, Poncher HG. 1941. malnourished Guatemalan children. American
Developmental pattern of the child as reflected Journal of Physical Anthropology 92:37–51.
in the calcification pattern of the teeth. Mayhall JT. 1970. The effect of culture change
American Journal of Diseases of Children upon the Eskimo dentition. Arctic
62:33–67. Anthropology 7:117–121.
Massola A. 1963. The parrying shields of south- Mayhall JT. 1977. The oral health of a Canadian
east Australia. Proceedings of the Royal Society Inuit community: an anthropological
of Victoria 76:227–232. approach. Journal of Dental Research 56
Mata-Míguez J, Overhotzer L, Rodríguez-Alegría (Special Issue C):C55–C60.
E, Kemp BM, Bolnick DA. 2012. The genetic Maynard Smith J, Savage RJG. 1959. The
impact of Aztec imperialism: ancient mechanics of mammalian jaws. School Science
mitochondrial DNA evidence from Xaltocan, Review 40:289–301.
Mexico. American Journal of Physical Mays S. 1999. Linear and appositional long bone
Anthropology 149:504–516. growth in earlier human populations: a case
Matheson CD, Vernon KK, Lahti A, et al. 2009. study from Mediaeval England. In: Hoppa RD,
Molecular exploration of the first-century FitzGerald CM, editors. Human Growth in
Tomb of the Shroud in Akeldama, Jerusalem. the Past: Studies from Bones and Teeth.
PLoS ONE 4:e8319–e-8319. Cambridge, UK: Cambridge University Press;
Matos V, Santos AL. 2006. On the trail of p. 290–312.
pulmonary tuberculosis based on rib lesions: Mays S. 2006a. Spondylolysis, spondylolisthesis,
results from the Human Identified Skeletal and lumbo-sacral morphology in a Medieval
Collection from the Museu Bocage (Lisbon, English skeletal population. American
Portugal). American Journal of Physical Journal of Physical Anthropology
Anthropology 130:190–200. 131:352–362.
Matsumura H. 2006. The population history of Mays SA. 2006b. Age-related cortical bone loss in
Southeast Asia from morphometric analyses of women in a 3rd–4th century AD population
human skeletal and dental remains. In: from England. American Journal of Physical
Oxenham M, Tayles N, editors. Bioarchaeology Anthropology 129:518–528.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
522 References

Mays S. 2008. A likely case of scurvy from Early and beyond. In: Steckel RH, Rose JC, editors.
Bronze Age Britain. International Journal of The Backbone of History: Long-Term Trends in
Osteoarchaeology 18:178–187. Health and Nutrition in the Americas. New
Mays S, Brickley M, Ives R. 2006. Skeletal York, NY: Cambridge University Press;
manifestations of rickets in infants and young p. 94–124.
children in a historic population from England. McCarroll JR, Miller JM, Ritter MA. 1986. Lumbar
American Journal of Physical Anthropology spondylolysis and spondylolisthesis in college
129:362–374. football players. American Journal of Sports
Mays S, Brickley M, Ives R. 2009. Growth and Medicine 14:404–406.
vitamin D deficiency in a population from 19th McDonagh MS, Whiting PF, Wilson PM, et al.
century Birmingham, England. International 2000. Systematic review of water fluoridation.
Journal of Osteoarchaeology 19:406–415. British Medical Journal 321:855–859.
Mays S, Crane-Kramer G, Bayliss A. 2003. Two McDowell MA, Fryar CD, Ogden CL, Flegal KM.
probable cases of treponemal disease of 2008. Anthropometric Reference Data for
Medieval date from England. American Journal Children and Adults: United States,
of Physical Anthropology 120:133–143. 2003–2006. National Health Statistics
Mays S, Fysh E, Taylor GM. 2002. Investigation of Reports, 10; p. 1–11.
the link between visceral surface rib lesions McEwan BG, editor. 1993. The Spanish Missions
and tuberculosis in a Medieval skeletal series of La Florida. Gainesville, FL: University Press
from England using ancient DNA. American of Florida.
Journal of Physical Anthropology 119:27–36. McGregor IA, Thomason AM, Billewicz WZ. 1968.
Mays S, Ives R, Brickley M. 2009. The effects of The development of primary teeth in children
socioeconomic status on endochrondral and from a group of Gambian villages and critical
appositional bone growth, and acquisition of examination of its use for estimating age.
cortical bone in children from 19th century British Journal of Nutrition 222:307–314.
Birmingham, England. American Journal of McGuire RH. 1994. Do the right thing. In: Bray RL,
Physical Anthropology 140:410–416. Killion TW, editors. Reckoning with the Dead:
Mays S, Taylor GM. 2003. A first prehistoric case The Larsen Bay Repatriation and the
of tuberculosis from Britain. International Smithsonian Institution. Washington, DC:
Journal of Osteoarchaeology 13:189–196. Smithsonian Institution Press; p. 180–183.
Mays S, Taylor GM, Legge AJ, Young DB, Turner- McHenry H. 1968. Transverse lines in long bones
Walker G. 2001. Paleopathological and of prehistoric California Indians. American
biomolecular study of tuberculosis in a Medieval Journal of Physical Anthropology 29:1–18.
skeletal collection from England. American McIlvaine BK, Schepartz LA, Larsen CS, Sciulli,
Journal of Physical Anthropology 114:298–311. PW. 2014. Evidence for long-term migration
Mays SA. 1996. Healed limb amputations in on the Balkan peninsula using dental and
human osteology and their causes: a case study cranial nonmetric data: early interaction
from Ipswich, UK. International Journal of between Corinth (Greece) and its colony at
Osteoarchaeology 6:101–113. Apollonia (Albania). American Journal of
McArthur M. 1960. Food consumption and dietary Physical Anthropology 153:236–248.
levels of groups of Aborigines living on McKeag DB. 1992. The relationship of
naturally occurring foods. In: Mountford CP, osteoarthritis and exercise. Clinics in Sports
editor. Records of the American-Australian Medicine 11:471–487.
Expedition to Arnhem Land, 1948. Volume II: McKeown AH, Jantz RL. 2005. A comparison of
Anthropology and Nutrition. Melbourne, craniometric and coordinate data for biological
Australia: Melbourn University Press. distance studies. In: Slice DE, editor. Modern
McCaa R. 2002. Paleodemography of the Morphometrics in Physical Anthropology. New
Americas: from ancient times to colonialism York, NY: Kluwer Academic Press; p. 215–230.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 523

McKern TW, Stewart TD. 1957. Skeletal Changes Pleistocene and early Holocene. In: Pinhasi R,
in Young American Males. Technical Report, Stock JT, editors. Human Bioarchaeology of the
No. EP-45. Natick, MA: U.S. Quartermaster Transition to Agriculture. Chichester, UK:
Research and Development Center. Wiley-Blackwell; p. 153–176.
McKinley JI. 1993. A decapitation from the Meiklejohn C, Baldwin JH, Schentag CT. 1988.
Romano-British cemetery at Baldock, Caries as a probable dietary marker in the
Hertfordshire. International Journal of western European Mesolithic. In: Kennedy BV,
Osteoarchaeology 3:41–44. LeMoine GM, editors. Diet and Subsistence:
McLean FC, Urist MR. 1968. Bone: Fundamentals Current Archaeological Perspectives.
of the Physiology of Skeletal Tissue, Third Chacmool, Archaeological Association of the
Edition. Chicago, IL: University of Chicago University of Calgary, Proceedings of the 19th
Press. Annual Conference; p. 273–279.
McMichael AJ. 2010. Paleoclimate and bubonic Meiklejohn C, Schentag C, Venema A, Key P.
plague: a forewarning of future risk? BMC 1984. Socioeconomic change and patterns of
Biology 8:108;www.biomedcentral.com/1741- pathology and variation in the Mesolithic and
7007/8/108. Neolithic of western Europe: some suggestions.
McMurray RG. 1995. Effects of physical activity In: Cohen MN, Armelagos GJ, editors.
on bone. In: Anderson JJB, Garner SC, editors. Paleopathology at the Origins of Agriculture.
Calcium and Phosphorus in Health and Orlando, FL: Academic Press; p. 75–100.
Disease. Boca Raton, FL: CRC Press; Meiklejohn C, Wyman JM, Jacobs K, Jackes MK.
p. 301–317. 1997. Issues in the archeological demography
McNeill WH. 1998. Plagues and Peoples. Garden of the agricultural transition in western and
City, NY: Anchor Press. northern Europe: a view from the Mesolithic.
McPhee IB, O’Brien JP. 1980. Scoliosis in In: Paine R, editor. Integrating Archaeological
symptomatic spondylolisthesis. Journal of Demography: Multidisciplinary Approaches to
Bone and Joint Surgery 62-B:155–157. Prehistoric Populations. Southern Illinois
McWhirr A, Viner L, Wells C, editors. 1982. University at Carbondale, Center for
Cirencester Excavations II: Romano-British Archaeological Investigations, Occasional
Cemeteries at Cirencester. Cirencester, UK: Paper, No. 24; p. 311–326.
Corinium Museum. Meiklejohn C, Wyman JM, Schentag CT. 1992.
Meade JB. 1989. The adaptation of bone to Caries and attrition: dependent or independent
mechanical stress: experimentation and variables? International Journal of
current concepts. In: Cowin SC, editor. Bone Anthropology 7:17–22.
Mechanics. Boca Raton, FL: CRC Press; Meiklejohn C, Zvelebil M. 1991. Health status of
p. 211–251. European populations at the agricultural
Medaglia CC, Little EA, Schoeninger MJ. 1990. transition and the implications for the adoption
Late Woodland diet on Nantucket Island: a of farming. In: Bush H, Zvelebil M, editors.
study using stable isotope ratios. Bulletin of the Health in Past Societies. British Archaeological
Massachusetts Archaeological Society Reports, International Series, No. 567;
51:49–60. p. 129–145.
Meehan B. 1977. Hunters by the seashore. Journal Meindl RS, Russell KF. 1998. Recent advances in
of Human Evolution 6:363–370. method and theory in paleodemography.
Meggitt M. 1977.Blood is their Argument: Warfare Annual Review of Anthropology 27:375–399.
among the Mae Enga Tribesmen of the New Melanson KJ. 2007. Nutrition review: diet,
Guinea Highlands. Palo Alto, CA: Mayfield nutrition, and osteoarthritis. American Journal
Publishing. of Lifestyle Medicine 1:260–263.
Meiklejohn C, Babb J. 2011. Long bone length, Melbye J, Fairgrieve SI. 1994. A massacre and
stature and time in the European late possible cannibalism in the Canadian Arctic:

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
524 References

new evidence from the Saunaktuk site Archaeological Survey of Canada, Mercury
(NgTn-1). Arctic Anthropology 31:57–77. Series Paper, No. 119.
Mellanby M. 1934. Diet and the Teeth: An Merbs CF. 1989a. Trauma. In: İşcan MY, Kennedy
Experimental Study, Part III. The Effect of Diet KAR, editors. Reconstruction of Life from the
on Dental Structure and Disease in Man. Skeleton. New York, NY: Liss; p. 161–189.
Medical Research Council Special Report Merbs CF. 1989b. Spondylolysis: its nature and
Series, No. 191. anthropological significance. International
Mellquist C, Sandberg T. 1939. Odontological Journal of Anthropology 4:163–169.
studies of about 1,400 Mediaeval skulls from Merbs CF. 1992. A new world of infectious disease.
Halland and Scania in Sweden and from the Yearbook of Physical Anthropology 35:3–42.
Norse colony in Greenland, and a contribution Merbs CF. 1995. Incomplete spondylolysis and
to the knowledge of their anthropology. healing: a study of ancient Eskimo skeletons.
Odontologisk Tidskrift 38:1–83. Spine 20:2328–2334.
Meltzer DJ. 1993. Pleistocene peopling of the Merbs CF. 2002. Asymmetrical spondylolysis.
Americas. Evolutionary Anthropology American Journal of Physical Anthropology
1:157–169. 119:156–174.
Meltzer DJ. 1995. Clocking the first Americans. Merbs CF, Euler RC. 1985. Atlanto-occipital fusion
Annual Review of Anthropology 24:21–45. and spondylolisthesis in an Anasazi skeleton
Meltzer DJ. 2009. First Peoples in a New World: from Bright Angel Ruin, Grand Canyon
Colonizing Ice Age America. Berkeley, CA: National Park, Arizona. American Journal of
University of California Press. Physical Anthropology 67:381–391.
Meng Y, Zhang H-Q, Pan F, et al. 2011. Prevalence Merbs CF, Miller RJ, editors. 1985. Health and
of dental caries and tooth wear in a Neolithic Disease in the Prehistoric Southwest. Arizona
population (6700–5600 years BP) from State University, Anthropological Research
northern China. Archives of Oral Biology Papers, No. 34.
56:1424–1435. Merchant VL, Ubelaker DH. 1977. Skeletal growth
Mensforth RP. 1981. Growth velocity and of the protohistoric Arikara. American Journal
chondroblastic stability as major factors of Physical Anthropology 46:61–72.
related to the pathogenesis and Merriwether DA, Rothhammer F, Ferrell RE. 1994.
epidemiological distribution of growth arrest Genetic variation in the New World: ancient
lines. American Journal of Physical teeth, bone, and tissue as sources of DNA.
Anthropology 54:253. Experientia 50:592–601.
Mensforth RP. 1985. Relative tibia long bone Merriwether DA, Rothhammer F, Ferrell RE. 1995.
growth in the Libben and Bt-5 prehistoric Distribution of the four founding lineage
skeletal populations. American Journal of haplotypes in Native Americans suggests a
Physical Anthropology 68:247–262. single wave of migration for the New World.
Mensforth RP, Lovejoy CO, Lallo JW, Armelagos American Journal of Physical Anthropology
GJ. 1978. The role of constitution factors, diet, 98:411–430.
and infectious disease in the etiology of porotic Metcalfe J, White CD, Longstaffe FJ. 2009.
hyperostosis and periosteal reactions in Hierarchies and heterarchies of food
prehistoric infants and children. Medical consumption: stable isotope evidence from
Anthropology 2:1–59. Chau Hiix and San Pedro, Belize. Latin
Merbs CF. 1980. The pathology of a La Jollan American Antiquity 20:15–36.
skeleton from Punta Minitas, Baja, California. Miao W, Tao W, Congcang Z, Wu L, Changsui W.
Pacific Coast Archaeological Society Quarterly 2013. Dental wear and oral health as indicators
16(4):37–43. of diet among the early Qin people. In:
Merbs CF. 1983. Patterns of Activity-Induced Pechenkina K, Oxenham M, editors.
Pathology in a Canadian Inuit Population. Bioarchaeology of East Asia: Movement,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 525

Contact, Health. Gainesville, FL: University Mills JO. 1992. Beyond nutrition: antibodies
Press of Florida; p. 265–287. produced through grain storage practices, their
Micozzi MS, Kelley MA. 1985. Evidence for recognition and implications for the Egyptian
pre-Columbian tuberculosis at the Point of Predynastic. In: Friedman R, Adams B, editors.
Pines site, Arizona: skeletal pathology in the The Followers of Horus: Studies Dedicated to
sacro-iliac region. In: Merbs CF, Miller RJ, Michael Allen Hoffman. Oxford, UK: Oxbow
editors. Health and Disease in the Prehistoric Books; p. 27–35.
Southwest. Arizona State University Milner GR. 1982. Measuring Prehistoric Levels of
Anthropological Research Papers, No. 34; Health: A Study of Mississippian Period
p. 347–358. Skeletal Remains from the American Bottom,
Mihesuah DA, editor. 2000. Repatriation Reader: Illinois. PhD Dissertation, Northwestern
Who Owns American Indian Remains? Lincoln, University, Evanston, IL.
NE: University of Nebraska Press. Milner GR. 1983. The East St. Louis Stone Quarry
Miles AEW. 1963. The dentition in the assessment Site Cemetery. American Bottom Archaeology,
of individual age in skeletal material. In: FAI-270 Site Reports, 1. Urbana, IL: University
Brothwell DR, editor. Dental Anthropology. of Illinois Press.
New York, NY: Pergamon Press; Milner GR. 1984. Dental caries in the permanent
p. 191–209. dentition of a Mississippian period population
Miles JS. 1975. Orthopedic Problems of the from the American Midwest. Collegium
Wetherill Mesa Populations, Mesa Verde Antropologicum 8:77–91.
National Park, CO. National Park Service, Milner GR. 1990. The late prehistoric Cahokia
Publications in Archaeology, No. 7G. cultural system of the Mississippi River valley:
Miller E. 1994. Evidence for prehistoric scalping in foundations, florescence, and fragmentation.
northeastern Nebraska. Plains Anthropologist Journal of World Prehistory 4:1–43.
39:211–219. Milner GR. 1991. Health and cultural change in
Miller EA. 1995. Refuse to be Ill: European Contact the late prehistoric American Bottom, Illinois.
and Aboriginal Health in Northeastern In: Powell ML, Bridges PS, Mires AMW, editors.
Nebraska. PhD dissertation, Arizona State What Mean These Bones? Studies in
University, Tempe, AZ. Southeastern Bioarchaeology. Tuscaloosa, AL:
Miller GJ, Piotrowski G. 1977. Geometric University of Alabama Press; p. 52–69.
properties of paired human femurs. In: Grood Milner GR 1992. Disease and sociopolitical
ES, Smith CR, editors. Advances in systems in late prehistoric Illinois. In: Verano
Bioengineering. New York, NY: American JW, Ubelaker DH, editors. Disease and
Society for Testing and Materials; p. 73–74. Demography in the Americas. Washington, DC:
Miller RJ. 1985. Lateral epicondylitis in a Smithsonian Institution Press; p. 103–116.
prehistoric central Arizona Indian population Milner GR. 1995. An osteological perspective on
from Nuvakwewtaqa (Chavez Pass). In: Merbs prehistoric warfare. In: Beck LA, editor.
CF, Miller RJ, editors. Health and Disease in the Regional Approaches to Mortuary Analysis.
Prehistoric Southwest. Arizona State New York, NY: Plenum Press; p. 221–244.
University Anthropological Research Papers, Milner GR. 1999. Warfare in prehistoric and early
34; p. 391–400. historic eastern North America. Journal of
Miller-Shaivitz P, İşcan MY. 1991. The prehistoric Archaeological Research 7:105–151.
people of Fort Center: physical and health Milner GR, Anderson E, Smith VG. 1991. Warfare
characteristics. In: Powell ML, Bridges PS, in late prehistoric west-central Illinois.
Mires AMW, editors. What Mean These Bones? American Antiquity 56:581–603.
Studies in Southeastern Bioarchaeology. Milner GR, Boldsen JL. 2012a. Skeletal age
Tuscaloosa, AL: University of Alabama Press; estimation: where we are and where we should
p. 131–147. go. In: Dirkmaat DC, editor. A Companion

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
526 References

Forensic Anthropology. Chichester, UK: Minagawa M, Wada E. 1984. Stepwise enrichment


Wiley-Blackwell; p. 224–238. of 15N along food chains: further evidence and
Milner GR, Boldsen JL. 2012b. Transition analysis: the relation between δ15N and animal age.
a validation study with known-age modern Geochimica et Cosmochimica Acta
American skeletons. American Journal of 48:1135–1140.
Physical Anthropology 148:98–110. Minns RJ, Bremble GR, Campbell J. 1975. The
Milner GR, Boldsen JL. 2012c. Estimating age and geometric properties of the human tibia.
sex from the skeleton, a paleopathological Journal of Biomechanics 8:253–255.
perspective. In: Grauer AL, editor. Minozzi S, Manzi G, Ricci F, di Lernia S,
A Companion to Paleopathology. Chichester, Borgognini Trali SM. 2003. Nonalimentary
UK: Wiley-Blackwell; p. 268–284. tooth use in prehistory: an example from early
Milner GR, Chaplin G, Zavodny E. 2013. Conflict Holocene in central Sahara (Uan Muhuggiag,
and societal change in late prehistoric Eastern Tadrart Acacus, Libya). American Journal of
North America. Evolutionary Anthropology Physical Anthropology 120:225–232.
22:96–102. Miritoiu N. 1992. Porotic hyperostosis in the Free
Milner GR, Humpf DA, Harpending HC. 1989. Dacians’ necropolis at Poisnesti (Vaslui
Pattern matching of age at death distributions County): congenital hemolytic anemia or iron
in paleodemographic analysis. American deficiency anemia? Preliminary
Journal of Physical Anthropology 80:49–58. considerations. Annuaire Roumain
Milner GR, Larsen CS. 1991. Teeth as artifacts of d’Anthropologie 29:3–12.
human behavior: intentional mutilation and Mitchell PD. 2003. Pre-Columbian treponemal
accidental modification. In: Kelley MA, Larsen disease from 14th century AD Safed, Israel, and
CS, editors. Advances in Dental Anthropology. implications for the Medieval eastern
New York, NY: Wiley-Liss; p. 357–378. Mediterranean. American Journal of Physical
Milner GR, Larsen CS, Hutchinson DL, Williamson Anthropology 121:117–124.
M., Humpf DA. 2000. Conquistadors, excavators, Mitchell PD, Nagar Y, Ellenblum R. 2006. Weapon
or rodents: what damaged the King site injuries in the 12th century crusader garrison
skeletons? American Antiquity 65:355–363. of Vadum Iacob castle, Galilee. International
Milner GR, Smith VG. 1989. Carnivore alteration Journal of Osteoarchaeology 16:145–155.
of human bone from a late prehistoric site in Mittler DM, Van Gerven DP. 1994. Developmental,
Illinois. American Journal of Physical diachronic, and demographic analysis of cribra
Anthropology 79:43–49. orbitalia in the Medieval Christian populations
Milner GR, Smith VG. 1990. Oneota human of Kulubnarti. American Journal of Physical
skeletal remains. In: Santure SK, Harn AD, Anthropology 93:287–297.
Esarey D, editors. Archaeological Investigations Miyamoto A, Shigematsu T, Fukunaga T, et al.
at the Morton Village and Norris Farms 35 1998. Medical baseline data collection on bone
Cemetery. Illinois State Museum Reports of change with space flight. Bone 22:79S–82S.
Investigations, No. 45; p. 111–148. Mizoguchi Y. 1985. Shovelling: A Statistical
Milner GR, Wood JW, Boldsen JL. 2008. Advances Analysis of Its Morphology. University
in paleodemography. In: Katzenberg MA, Museum, University of Tokyo, Bulletin, No. 26.
Saunders SR, editors. Biological Anthropology Mizoguchi Y. 1986. Correlated asymmetries
of the Human Skeleton, Second Edition. detected in the tooth crown diameters of
Chichester, UK: Wiley-Blackwell; p. 561–600. human permanent teeth. Bulletin of the
Minagawa M, Akazawa T. 1992. Dietary patterns National Science Museum, Tokyo Series
of Japanese Jomon hunter-gatherers: stable 12:24–45.
nitrogen and carbon isotope analyses of Mobley C. 1980. Demographic structure of Pecos
human bones. Pacific Northeast Asia Indians: a model based on life tables. American
Prehistory 1992; p.59–68. Antiquity 45:518–530.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 527

Moffat T. 2003. Diarrhea, respiratory infections, Molleson T. 1994. The eloquent bones of Abu
protozoan gastrointestinal parasites, and child Hureya. Scientific American 271(2):70–75.
growth in Kathmandu, Nepal. American Molleson T. 1995. Rates of ageing in the
Journal of Physical Anthropology 122:85–97. eighteenth century. In: Saunders SR, Herring A,
Moffat T, Galloway T. 2007. Adverse editors. Grave Reflections: Portraying the Past
environments: investigating local variation in through Cemetery Studies. Toronto, ON:
child growth. American Journal of Human Canadian Scholars’ Press; p. 199–222.
Biology 19:676–683. Molleson T. 2003. Body of evidence: museum
Moggi-Cecchi J, Pacciani E, Pinto-Cisternas J. collection, why they were brought together,
1994. Enamel hypoplasia and age at weaning their value today and public future. In: Grupe
in 19th-century Florence, Italy. American G, Peters J, editors. Decyphering Ancient
Journal of Physical Anthropology 93: Bones: The Research Potential of
299–306. Bioarchaeological Collections. Rahden,
Mokyr J, Ó Gráda C. 1994. The heights of the Germany: Verlag Leidorf GmbH; p. 17–28.
British and the Irish c. 1800–1815: evidence Molleson T, Cox M. 1993. The Spitalfields Project:
from recruits to the East India Company’s Volume 2. The Anthropology, The Middling
army. In: Komlos J, editor. Stature, Living Sort. Council for British Archaeology Report,
Standards, and Economic Development: Essays No. 86.
in Anthropometric History. Chicago, IL: Molleson T, Jones K. 1991. Dental evidence for
University of Chicago Press; p. 39–59. dietary changes at Abu Hureyra. Journal of
Møller AP, Swaddle JP. 1997. Asymmetry, Archaeological Science 24:455–468.
Developmental Stability, and Evolution. Molleson T, Jones K, Jones S. 1993. Dietary
Oxford, UK: Oxford University Press. change and the effects of food preparation on
Møller-Christensen V. 1958. Bogen om Æbelholt microwear patterns in the Late Neolithic of Abu
Kloster. Copenhagen, Denmark: Dansk Hureyra, northern Syria. Journal of Human
Videnskabs Forlag. Evolution 24:455–468.
Møller-Christensen V. 1961. Bone Changes in Molnar P. 2006. Tracing prehistoric activities:
Leprosy. Copenhagen, Denmark: Munksgaard. musculoskeletal stress marker analysis of a
Møller-Christensen V. 1978. Leprosy Changes of Stone-age population on the island of Gotland
the Skull. Odense, Denmark: Odense University in the Baltic Sea. American Journal of Physical
Press. Anthropology 120:12–23.
Møller-Christensen V, Hughes DR. 1966. An early Molnar P. 2008. Dental wear and oral pathology:
case of leprosy from Nubia. Man 62:177–179. possible evidence and consequences of
Møller-Christensen V, Inkster RG. 1965. Cases of habitual use of teeth in a Swedish Neolithic
leprosy and syphilis in the osteological sample. American Journal of Physical
collection of the Department of Anatomy, Anthropology 136:423–431.
University of Edinburgh, with note on the skull Molnar P, Ahlstrom TP, Leden I. 2011.
of King Robert the Bruce. Danish Medical Osteoarthritis and activity: an analysis of the
Bulletin 12:11–18. relationship between eburnation,
Molleson T. 1989. Seed preparation in the musculoskeletal stress markers (MSM) and age
Mesolithic: the osteological evidence. in two Neolithic hunter-gatherer populations
Antiquity 63:356–362. from Gotland, Sweden. International Journal of
Molleson T. 1992. Mortality patterns in the Osteoarchaeology 21:283–291.
Romano-British cemetery at Poundbury Camp Molnar S. 1971. Human tooth wear, tooth
near Dorchester. In: Bassett S, editor. Death in function and cultural variability. American
Towns: Urban Responses to the Dying and the Journal of Physical Anthropology 34:175–189.
Dead, 100–1600. Leicester, UK: Leicester Molnar S. 1972. Tooth wear and culture: a survey
University Press; p. 43–55. of tooth functions among some prehistoric

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
528 References

populations. Current Anthropology Montgomery J, Evans J, Chenery S, Pashley V,


13:511–526. Killgrove K. 2010. “Gleaming, white and
Molnar S, Hildebolt C. 1987. Geochemical deadly”: using lead to track human exposure
influences in dental disease. Anthropologiai and geographic origins in the Roman period in
Közlemenyek 31:3–16. Britain. Journal of Roman Archaeology
Molnar S, McKee JK, Molnar IM, Przybeck TR. Supplement 78:199–226
1983. Tooth wear rates among contemporary Montgomery J, Knüsel CJ, Tucker K. 2011.
Australian Aborigines. Journal of Dental Identifying from origins of decapitated male
Research 62:562–565. skeletons from 3 Driffield Terrace, York,
Molnar S, Molnar IM. 1985a. The incidence of through isotope analysis. In: Bonogofsky M,
enamel hypoplasia among the Krapina editor. The Bioarchaeology of the Human Head:
Neandertals. American Anthropologist Decapitation, Decoration, and Deformation.
87:536–549. Gainesville, FL: University Press of Florida;
Molnar S, Molnar I. 1985b. Observations of dental p. 141–178.
diseases among prehistoric populations of Montgomery RT, Perry M. 2012. The social and
Hungary. American Journal of Physical cultural implications of violence at Qasr
Anthropology 67:51–63. Hallabat. In: Martin DL, Harrod RP, Pérez VR,
Molnar S, Molnar IM. 1990. Dental arch shape and editors. The Bioarchaeology of Violence.
tooth wear variability. American Journal of Gainesville, FL: University Press of Florida;
Physical Anthropology 82:385–395. p. 83–110.
Molnar S, Richards L, McKee J, Molnar I. 1989. Moore JA, Swedlund AC, Armelagos GJ. 1975. The
Tooth wear in Australian Aboriginal use of life tables in paleodemography. In:
populations from the River Murray valley. Swedlund AC, editor. Population Studies in
American Journal of Physical Anthropology Archaeology and Biological Anthropology.
79:185–196. Memoirs of the Society for American
Molto JE. 2002. Leprosy in Roman period Archaeology, No. 30; p. 57–70.
skeletons from Kellis 2, Dakhleh, Egypt. In: Moore KP, Thorp S, Van Gerven DP. 1986. Pattern
Roberts CA, Lewis ME, Manchester K, editors. of dental eruption, skeletal maturation and
The Past and Present of Leprosy: stress in a Medieval population from Sudanese
Archaeological, Historical, Palaeopathological Nubia. Human Evolution 1:325–330.
and Clinical Approaches. British Moore MR. 1986. Lead in humans. In: Lansdown
Archaeological Reports, International Series, R, Yule W, editors. Lead Toxicity: History and
No. 1054; p. 179–192. Environmental Impact. Baltimore, MD: Johns
Monahan EI. 1995. Bioarchaeological analysis of Hopkins University Press; p. 54–95.
the mortuary practices at the Broad Reach site Moore RI. 1987. The Formation of a Persecuting
(31CR218), coastal North Carolina. Southern Society: Power and Deviance in Western
Indian Studies 44:37–69. Europe, 950–1250. New York, NY: Basil
Montagu A. 1978. Learning Non-Aggression: The Blackwell.
Experience of Non-Literate Societies. New Moore WJ. 1965. Masticatory function and skull
York, NY: Oxford University Press. growth. Journal of Zoology 146:123–131.
Montgomery J, Evans JA. 2006. Immigrants on the Moore WJ. 1967. Muscular function and skull
Isle of Lewis: combining traditional funerary growth in the laboratory rat (Rattus
and modern isotope evidence to investigate norvegicus). Journal of Zoology 152:287–296.
social differentiation, migration and dietary Moore WJ. 1973. An experimental study of the
change in the Outer Hebrides of Scotland. In: functional components of growth in the rat
Gowland R, Knüsel C, editors. Social mandible. Acta Anatomica 85:378–385.
Archaeology of Funerary Remains. Oxford, UK: Moore WJ, Corbett ME. 1971. The distribution of
Oxbow Books; p. 122–142. dental caries in ancient British populations:

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 529

Anglo-Saxon period. Caries Research Archaeology and Ethnology, Harvard


5:151–161. University, 85; p. 27–42.
Moore WJ, Corbett ME. 1973. The distribution of Morimoto I. 1984. Comment. Current
dental caries in ancient British populations: Anthropology 25:322.
Iron Age, Romans, British, and Medieval Morris AG. 1992. The Skeletons of Contact:
periods. Caries Research 7:139–151. A Study of Protohistoric Burials from the
Moore WJ, Corbett ME. 1975. The distribution of Lower Orange River Valley, South Africa.
dental caries in ancient British populations. Johannesburg, South Africa: Witwatersrand
Caries Research 9:163–174. University Press.
Moore WJ, Lavelle CLB, Spence TF. 1968. Changes Morris AG. 2009. The politics of old bones in 21st
in the size and shape of the human mandible. century South Africa. The Digging Stick
British Dental Journal 125:163–169. 26:5–9.
Moorrees CFA. 1957. The Aleut Dentition: Morris NT. 1981. The occurrence of mandibular
A Correlative Study of Dental Characteristics in torus. In: Hayes AC, editor. Contributions to
an Eskimoid People. Cambridge, MA: Harvard Gran Quivira Archeology. National Park
University Press. Service, Publications in Archeology, No. 17;
Moorrees CFA, Fanning EA, Hunt EE. 1963. Age p. 123–127.
variation of formation stages for ten Morse D. 1967. Tuberculosis. In: Brothwell D,
permanent teeth. Journal of Dental Research Sandison AT, editors. Diseases in Antiquity.
42:1490–1502. Springfield, IL: Charles C. Thomas; p. 249–271.
Moorrees CFA, Kent RL. 1981. Interrelations in the Morton SG. 1844. Crania Aegyptiaca.
timing of root formation and tooth emergence. Transactions of the American Philosophical
Proceedings of the Finnish Dental Society Society, 9.
77:113–117. Moseley JE. 1974. Skeletal changes in the anemias.
Moraga M, Santoro CM, Standen VG, Carvallo P, Seminars in Roentgenology 9:169–184.
Rothhammer F. 2005. Microevolution in Moskowitz RW, Kelly MA, Lewallen DG. 2004.
prehistoric Andean populations: chronologic Understanding osteoarthritis of the knee –
mtDNA variation in the desert valleys of causes and effects. American Journal of
northern Chile. American Journal of Physical Orthopaedics 33(2S):5–9.
Anthropology 127:170–181. Motarjemi Y, Kaferstein F, Moy G, Quevedo F.
Morant GM. 1925. A study of Egyptian craniology 1993. Contaminated weaning food: a major
from prehistory to Roman times. Biometrika risk factor for diarrhea and associated
17:1–52. malnutrition. Bulletin of the World Health
Morell V. 1995. Who owns the past? Science Organization 71:79–92.
268:1424–1426. Mukherjee R, Rao CR, Trevor JC. 1955. The
Moreton RD. 1966. Spondylolysis. Journal of Ancient Inhabitants of Jebel Moya (Sudan).
the American Medical Association Cambridge, UK: Cambridge University Press.
195:671–674. Müldner G, Richards MP. 2007a. Stable isotope
Morey ER, Baylink DJ. 1978. Inhibition of bone evidence for 1500 years of human diet at the
formation during spaceflight. Science city of York, UK. American Journal of Physical
201:1138–1141. Anthropology 133:682–697.
Morgan ME. 2010. A reassessment of the human Müldner G, Richards MP. 2007b. Diet and diversity
remains from the Upper Pecos valley formerly at later Medieval Fishergate: the isotopic
curated at the Peabody Museum of evidence. American Journal of Physical
Archaeology and Ethnology, Harvard Anthropology 134:162–174.
University. In: Morgan ME, editor. Pecos Mulhern DM. 1996. The effects of biology and
Pueblo Revisited: The Biological and Social environment on patterns of microscopic
Context. Papers of the Peabody Museum of histologic change in human bone. American

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
530 References

Journal of Physical Anthropology Supplement Archaeological Reports, International Series,


22:173–174. No. 1152.
Mulhern DM. 2000. Rib remodeling dynamics in a Murphy EM. 2007. Human osteoarchaeology in
skeletal population from Kulubnarti, Nubia. Ireland. In: Murphy EM, Whitehouse NJ,
American Journal of Physical Anthropology editors. Environmental Archaeology in Ireland.
111:519–530. Oxford, UK: Oxbow Books; p. 48–76.
Mulhern DM, Van Gerven DP. 1997. Patterns of Murphy EM, Chistov YK, Hopkins R, et al. 2009.
femoral bone remodeling dynamics in a Tuberculosis among Iron Age individuals from
medieval Nubian population. American Tyva, South Siberia: palaeopathological and
Journal of Physical Anthropology biomolecular findings. Journal of
104:123–146. Archaeological Science 36:2029–2038.
Müller R, Roberts CA, Brown TA. 2014. Murphy M, Gaither C, Goycochea E, Verano JW,
Biomolecular identification of ancient Cock G. 2010. Violence and weapon-related
Mycobacterium tuberculosis complex DNA in trauma at Puruchuco-Huaquerones, Peru.
human remains from Britain and continental American Journal of Physical Anthropology
Europe. American Journal of Physical 142:636–649.
Anthropology 153:178–189. Murray KA. 1989. Bioarchaeology of the Parkin
Mulligan C, Hunley K, Cole S, Long J. 2004. site. Pittsburgh Undergraduate Review 6:1–20.
Population genetics, history, and health Murray ML, Schoeninger MJ. 1988. Diet, status,
patterns in Native Americans. Annual Review and complex social structure in Iron Age
of Genomics and Human Genetics 5:295–315. central Europe: some contributions of bone
Mummert A, Esche E, Robinson J, Armelagos GJ. chemistry. In: Gibson DB, Geselowitz MN,
2011. Stature and robusticity during the editors. Tribe and Polity in Late Prehistoric
agricultural transition: Evidence from the Europe. New York, NY: Plenum Press;
bioarchaeological record. Economics and p. 155–176.
Human Biology 9:284–301. Mushrif-Tripathy V, Walimbe SR. 2006. Human
Mummery JR. 1870. On the relations which dental Skeletal Remains from Chalcolithic Nevasa:
caries, as discovered among the ancient Osteobiographic Analysis. British
inhabitants of Britain and amongst existing Archaeological Reports, International Series,
aboriginal races, may be supposed to hold to No. 1476.
their food and social conditions. Transactions Myers F. 1986. Pintupi Country, Pintupi Self:
of the Odontological Society of Great Britain Sentiment, Place, and Politics among Western
2:7–80. Desert Aborigines. Washington, DC:
Murchison MA, Riopelle AJ, Owsley DW. 1988. Smithsonian Institution Press.
Dental development in protein-deprived infant Nagaoka T, Hirata K. 2007. Reconstruction of
rhesus monkeys. Human Biology 60:383–394. paleodemographic characteristics from skeletal
Murphy E, Manchester K. 2002. Evidence for age at death distributions: perspectives from
leprosy in Medieval Ireland. In: Roberts CA, Hitotsubashi, Japan. American Journal of
Lewis ME, Manchester K, editors. The Past and Physical Anthropology 134:301–311.
Present of Leprosy: Archaeological, Historical, Nagaoka T, Hirata K, Yokota E, Matsu’ura S. 2006.
Palaeopathological and Clinical Approaches. Paleodemography of a Medieval population in
Oxford, UK: British Archaeological Reports, Japan: analysis of human skeletal remains
International Series; p. 193–200. from the Yuigahama-minami site. American
Murphy EM. 2002. Human osteoarchaeology in Journal of Physical Anthropology 131:1–14.
Ireland: past, present and future. Antiquity Nagurka ML, Hayes WC. 1980. An interactive
76:512–517. graphics package for calculating cross-
Murphy EM. 2003. Iron Age Archaeology and sectional properties of complex shapes. Journal
Trauma from Aymyrlyg, South Siberia. British of Biomechanics 13:59–64.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 531

Nakahashi T. 1993. Temporal craniometric Santa Cruz Island, California. American


changes from the Jomon to the Modern period Journal of Physical Anthropology 88:135–144.
in western Japan. American Journal of Physical Nelson GC, Lukacs JR, Yule P. 1999. Dates, caries
Anthropology 90:409–425. and early tooth loss during the Iron Age of
Nakamura R, Ono Y, Horii E, Tsunoda K, Takeuchi Oman. American Journal of Physical
Y. 1993. The aetiological significance of work- Anthropology 108:333–343.
load in the development of osteoarthritis of the Nelson ML, Dinardo A, Hochberg J, Armelagos GJ.
distal interphalangeal joint. Journal of Hand 2010. Mass spectroscopic characterization of
Surgery 18B:540–542. tetracycline in the skeletal remains of an
Nakashima T, Matsuno K, Matsushita M, ancient population from Sudanese Nubia
Matsushita T. 2011. Severe lead contamination 350–550 CE. American Journal of Physical
among children of Samurai families in Edo Anthropology 143:151–154.
period Japan. Journal of Archaeological Nelson PA, Sauer NJ. 1984. An evaluation of
Science 38:23–28. postdepositional changes in the trace element
Nanci A, editor. 2013. Ten Cate’s Oral Histology: content of human bone. American Antiquity
Development, Structure, and Function, Eighth 49:141–147.
Edition. Amsterdam, the Netherlands: Elsevier. Neumann GK. 1940. Evidence for the antiquity of
Nass GG. 1982. Dental asymmetry as an indicator scalping from central Illinois. American
of developmental stress in a free-ranging troop Antiquity 4:287–289.
of Macaca fuscata. In: Kurtén B, editor. Teeth: Neumann GK. 1952. Archeology and race in the
Form, Function, and Evolution. New York, NY: American Indian. In: Griffin JB, editor.
Columbia University Press; p. 207–227. Archeology of the Eastern United States.
Nathan H. 1959. Spondylolysis: its anatomy and Chicago, IL: University of Chicago Press;
mechanisms of development. Journal of Bone p. 13–34.
and Joint Surgery 41-A:303–320. Neves WA, Barros AM, Costa MA. 1999. Incidence
Navas-Carretero S, Pérez-Granados AM, Sarriá B, and distribution of postcranial fractures in the
et al. 2008. Oily fish increases iron prehistoric population of San Pedro de
bioavailability of a phytate rich meal in young Atacama, northern Chile. American Journal of
iron defient women. Journal of the American Physical Anthropology 109:253–258.
College of Nutrition 27:96–101. Neves WA, Gonzalez-Jose R, Hubbe M, et al. 2004.
Nawrocki SP. 2010. The nature and sources of Early Holocene humans skeletal remains from
error in the estimation of age at death from the Cerca Grande, Lagoa Santa, Central Brazil, and
skeleton. In: Latham KE, Finnegan M, editors. the origin of the first Americans. World
Age Estimation of the Human Skeleton. Archaeology 36:479–501.
Springfield, IL: Charles C Thomas; p. 79–101. Neves WA, Hubbe M. 2005. Cranial morphology of
Nelson BK, DeNiro MJ, Schoeninger MJ, DePaolo early Americans from Lagoa Santa, Brazil:
DJ. 1986. Effects of diagenesis on strontium, implications for the settlement of the New
carbon, nitrogen, and oxygen concentration World. Proceedings of the National Academy of
and isotopic composition of bone. Geochimica Sciences 102:18309–18314.
et Cosmochimica Acta 50:1941–1949. Neves WA, Hubbe M, Correal G. 2007. Human
Nelson DA, Pettifor JM, Barondess DA, et al. 2004. skeletal remains from Sabana de Bogota,
Comparison of cross-sectional geometry of Colombia: a case of Paleoamerican
the proximal femur in white and black morphology late survival in South America?
women from Detroit and Johannesburg. American Journal of Physical Anthropology
Journal of Bone and Mineral Research 133:1080–1098.
19:560–565. Neves WA, Hubbe M, Okumura MMM, et al. 2005.
Nelson GC. 1992. Maxillary canine/third premolar A new early Holocene human skeleton from
transposition in a prehistoric population from Brazil: implications for the settlement of the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
532 References

New World. Journal of Human Evolution the onset of agriculture. American Journal of
48:403–414. Physical Anthropology 149:391–404.
Neves WA, Prous A, González-José R, Kipnis R, Nielsen OV. 1970. The Nubian Skeleton through
Powell J. 2003. Early Holocene human skeletal 4,000 Years. PhD Dissertation, University of
remains from Santana do Riacho, Brazil: Odense, Odense, Denmark.
implications for the settlement of the New Niinimäki S. 2012. The relationship between
World. Journal of Human Evolution 45:19–42. musculoskeletal stress markers and
Newbrun E. 1982. Sugar and dental caries: a review biomechanical properties of the humeral
of human studies. Science 217:418–423. diaphysis. American Journal of Physical
Newcomb WW Jr. 1950. A re-examination of the Anthropology 147:618–628.
causes of Plains warfare. American Nikiforuk G, Fraser D. 1984. Enhanced caries
Anthropologist 52:317–330. susceptibility associated with neonatal enamel
Newman J. 1995. How breast milk protects hypoplasia. Caries Research 18:185.
newborns. Scientific American 273:76–79. Nitsch EK, Humphrey LT, Hedges REM. 2011.
Newman MT. 1951. The sequence of Indian Using stable isotope analysis to examine the
physical types in South America. In: Laughlin effect of economic change in breastfeeding
WS, editor. The Physical Anthropology of the practices in Spitalfields, London, UK. American
American Indian. New York, NY: The Viking Journal of Physical Anthropology
Fund; p. 69–97. 146:619–628.
Newman MT. 1962. Evolutionary changes in body Nizel AE. 1973. Nutrition and oral problems.
size and head form in American Indians. World Review of Nutrition and Diet
American Anthropologist 64:237–257. 16:226–252.
Newman MT, Snow CE. 1942. Preliminary report Noback ML, Harvati K, Spoor F. 2011. Climate-
on the skeletal material from Pickwick Basin, related variation in the human nasal cavity.
Alabama. In: Webb WS, DeJarnette DL, editors. American Journal of Physical Anthropology
An Archaeological Survey of Pickwick Basin in 145:599–614.
the Adjacent Portions of the States of Alabama, Nordin C, Need AG, Morris HA. 1993. Metabolic
Mississippi, and Tennessee. Bureau of Bone and Stone Disease, Third Edition.
American Ethnology, Bulletin, No. 129; Edinburgh, UK: Churchill Livingstone.
p. 393–507. Nordin M, Frankel VH. 2001. Basic Biomechanics
Newton JS, Domett KM, O’Reilly DJW, Shewan L. of the Musculoskeletal System, Third Edition.
2013. Dental health in Iron Age Cambodia: Philadelphia, PA: Lippincott Williams &
temporal variations with rice agriculture. Wilkins.
International Journal of Paleopathology 3:1–10. Norr L. 1984. Prehistoric subsistence and health
Nichols DL, Crown PL, editors. 2008. Social status of coastal peoples from the Panamanian
Violence in the Prehispanic American isthmus or lower Central America. In: Cohen
Southwest. Tucson, AZ: University of Arizona MN, Armelagos GJ, editors. Paleopathology at
Press. the Origins of Agriculture. Orlando, FL:
Nicholson G. 1945. The two main diameters at the Academic Press; p. 463–490.
brim of the female pelvis. Journal of Anatomy Norr L. 1991. Nutritional Consequences of
79:131–135. Prehistoric Subsistence Strategies in Lower
Nickens PR. 1976. Stature reduction as an Central America. PhD Dissertation, University
adaptive response to food production in of Illinois, Urbana–Champaign, IL.
Mesoamerica. Journal of Archaeological Norr L. 2002. Stable isotope analysis and dietary
Science 3:31–41. inference. In: Hutchinson DL. Foraging,
Nicklisch N, Maixner F, Ganslmeier R,. 2012. Rib Farming, and Coastal Biocultural Adaptation
lesions in skeletons from early Neolithic sites in in Late Prehistoric North Carolina. Gainesville,
central Germany: on the trail of tuberculosis at FL: University Press of Florida; p. 178–205.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 533

Novak M, Šlaus M. 2011. Vertebral pathologies in editors. Advances in Human Palaeopathology.


two early modern period (16th–19th century) Chichester, UK: John Wiley & Sons, Ltd;
populations from Croatia. American Journal of p.283–307.
Physical Anthropology 145:270–281. Ogilvie MD, Curran BK, Trinkaus E. 1989.
Novak SA. 2000. Battle-related trauma. In: Fiorato Incidence and patterning of dental enamel
V, Boylston A, Knüsel C, editors. Blood Red hypoplasia among the Neandertals. American
Roses: The Archaeology of a Mass Grave from Journal of Physical Anthropology 79:25–41.
the Battle of Towton AD 1461. Oxford, UK: Ogilvie MD, Hilton CE. 2011. Cross-sectional
Oxbow Books; p. 90–102. geometry in the humeri of foragers and farmers
Novak SA. 2006. Beneath the facade: a skeletal from the Prehispanic American Southwest:
model of domestic violence. In: Gowland R, exploring patterns in the sexual division of
Knüsel C, editors. Social Archaeology of labor. American Journal of Physical
Funerary Remains. Oxford, UK: Oxbow Books; Anthropology 144:11–21.
p. 238–252. Okazaki K. 2004. A morphological study on the
Novak SA. 2014. How to say things with bodies: growth patterns of ancient people in northern
meaningful violence on an American frontier. Kyushu-Yamaguchi region, Japan.
In: Knüsel C, Smith MJ, editors. The Routledge Anthropological Science 112:219–234.
Handbook of the Bioarchaeology of Human Okazaki K, Pei-Ying T, Kuo-Shyan L. 2013. Sex
Conflict. London, UK: Routledge; p. 542–559. difference in oral disease of millet
Null CC, Blakey ML, Shujaa KJ, Rankin-Hill LM, agriculturalists from the Take-vatan lineage of
Carrington SHH. 2009. Osteological indicators the recent Bunan tribe of Taiwan.
of infectious disease and nutritional adequacy. Anthropological Science 121:105–113.
In: Blakey ML, Rankin-Hill LM, editors. The Okazaki K, Wei D, Zhu H. 2014. Variations in the
Skeletal Biology of the New York African Burial oral health of millet agriculturalists in the
Ground, Part I. Washington, DC: Howard northern “Great Wall” region of China from the
University Press; p. 169–198. Middle Neolithic to the Sixteen Kingdoms
Nuorala E. 1999. Tuberculosis on the 17th century period. Anthropological Science 121:
man-of-war Kronan. International Journal of 187–201.
Osteoarchaeology 9:344–348. O’Neill MC, Ruff CB. 2004. Estimating human long
Nystrom KC. 2011. Dental evidence of congenital bone cross-sectional geometric properties: a
syphilis in a 19th century cemetery from the comparison of noninvasive methods. Journal
mid-Hudson valley. International Journal of of Human Evolution 47:221–235.
Osteoarchaeology 21:371–378. Ophel IL. 1963. The fate of radiostrontium in a
Nystrom KC, Malcom CM. 2010. Sex-specific freshwater community. In: Schultz V, Klement
phenotypic variability and social organization AW, editors. Radioecology. London, UK:
in the Chiribaya of southern Peru. Latin Chapman and Hall; p. 213–216.
American Antiquity 21:375–397. Oranje P, Noriskin JN, Osborn TWB. 1935–37.
O’Connell BLH. 1983. Fluctuating Asymmetry as a The effect of diet upon dental caries in the
Measure of Developmental Stability in Illinois South African Bantu. South African Journal of
Woodland Populations. PhD dissertation, Medical Science 1–2:57–62.
Northwestern University, Evanston, IL. Organ JM, Teaford MF, Larsen CS. 2005. Dietary
O’Connell JF, Hawkes K, Blurton-Jones NG. 1992. inferences from dental occlusal microwear at
Patterns of distribution, site structure and Mission San Luis de Apalachee. American
assemblage composition of Hadza kill- Journal of Physical Anthropology
butchering sites. Journal of Archaeological 128:801–811.
Science 19:319–345. O’Rourke DH. 2010. Human molecular genetics:
Ogden A. 2008. Advances in the palaeopathology the DNA revolution and variation. In: Larsen
of teeth and jaws. In: Pinhasi R, Mays S, CS, editor. A Companion to Biological

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
534 References

Anthropology. Malden, MA: Wiley-Blackwell; Ortner DJ, Butler W, Cafarella J, Milligan L. 2001.
p. 88–103. Evidence of probable scurvy in subadults from
O’Rourke DH, Hayes MG, Carlyle SW. 2000. archeological sites in North America. American
Ancient DNA studies in physical anthropology. Journal of Physical Anthropology
Annual Review of Anthropology 29: 114:343–351.
217–242. Ortner DJ, Frohlich B, editors. 2008. The Early
O’Rourke DH, Parr RL, Carlyle SW. 1999. Bronze Age Tombs and Burials of Bab edh-
Molecular genetic variation in prehistoric Dhra’, Jordan. Lanham, MD: AltaMira.
inhabitants of the eastern Great Basin. In: Ortner DJ, Garofalo EM, Frohlich B. 2008. The
Hemphill BE, Larsen CS, editors. Prehistoric paleopathology of the EB IA and EB IB people.
Lifeways in the Great Basin Wetlands: In: Ortner DJ, Frohlich B, editors. The Early
Bioarchaeological Reconstruction and Bronze Age I Tombs and Burials of Bab edh-
Interpretation. Salt Lake City, UT: University of Dhra’, Jordan. Lanham, MD: Altamira;
Utah Press; p. 84–102. p. 263–280.
O’Rourke DH, Raff JA. 2010. The human genetic Ortner DJ, Kimmerle EH, Diez M. 1999. Probable
history of the Americas: the final frontier. evidence of scurvy in subadults from
Current Biology 20:R202–R207. archeological sites in Peru. American Journal
Orschiedt J, Haidle MN. 2012. Violence against the of Physical Anthropology 108:321–333.
living, violence against the dead on the human O’Shea J. 1984. Mortuary Variability: An
remains from Herxheim, Germany: evidence of Archaeological Investigation. Orlando, FL:
a crisis and mass cannibalism? In: Schulting R, Academic Press.
Fibiger L, editors. Sticks, Stones, & Broken O’Shea JM, Bridges PS. 1989. The Sargent site
Bones: Neolithic Violence in a European ossuary (25CU28), Custer County, Nebraska.
Perspective. Oxford, UK: Oxford University Plains Anthropologist 34:7–21.
Press; p. 121–137. Ossenberg N. 1970. The influence of artificial
Orschiedt J, Hauber A, Haidle MN, Alt KW, deformation on discontinuous morphological
Buitrago-Tellez CH. 2003. Survival of a traits. American Journal of Physical
multiple skull trauma: the case of an early Anthropology 33:357–372.
Neolithic individual from the LBK enclosure at Ossenberg NS. 1986. Isolate conservatism and
Herxheim (southwest Germany). International hybridization in the population history of
Journal of Osteoarchaeology 13:375–383. Japan: the evidence of nonmetric cranial traits.
Ortner DJ. 1968. Description and classification of In: Akazawa T, Aikens CM, editors. Prehistoric
degenerative bone changes in the distal joint Hunter-Gatherers in Japan: New Research
surfaces of the humerus. American Journal of Methods. Tokyo, Japan: University of Tokyo
Physical Anthropology 28:139–156. Press; p. 199–215.
Ortner DJ. 1979. Disease and mortality in the Early O’Sullivan EA, Williams SA, Wakefield RC, Cape
Bronze Age people of Bab edh-Dhra, Jordan. JE, Curzon MEJ. 1993. Prevalence and site
American Journal of Physical Anthropology characteristics of dental caries in primary
51:589–598. molar teeth from prehistoric times to the 18th
Ortner DJ. 1991. Theoretical and methodological century in England. Caries Research
issues in paleopathology. In: Ortner DJ, 27:147–153.
Aufderheide AC, editors. Human Ousley SD, Billeck WT, Hollinger RE. 2005. Federal
Paleopathology: Current Syntheses and Future repatriation legislation and the role of physical
Options. Washington, DC: Smithsonian anthropology in repatriation. Yearbook of
Institution Press; p. 5–11. Physical Anthropology 48:2–32.
Ortner DJ. 2003. Identification of Pathological Owen R. 1855. Of the anthropoid apes and their
Conditions in Human Skeletal Remains, Second relation to man. Proceedings of the Royal
Edition. San Diego, CA: Academic Press. Institution of Great Britain 58:26–41.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 535

Owsley DW. 1991. Temporal variation in femoral War of 1812. Toronto, ON: Dundurn Press;
cortical thickness of North American Plains p. 198–226.
Indians. In: Ortner DJ, Aufderheide AC, editors. Owsley DW, Morey DF, Turner WB. 1981.
Human Paleopathology: Current Syntheses and Inferring history from crania: biological
Future Options. Washington, DC: Smithsonian distance comparisons of Mill Creek and early
Institution Press; p. 105–110. Middle Missouri Tradition crania with Mandan
Owsley DW. 1994. Warfare in Coalescent tradition and Arikara populations samples. Plains
populations of the northern Plains. In: Owsley Anthropologist 26:301–310.
DW, Jantz RL, editors. Skeletal Biology in the Owsley DW, Orser CE, Mann RW, Moore-Jansen
Great Plains: Migration, Warfare, Health, and PH, Montgomery RL. 1987. Demography and
Subsistence. Washington, DC: Smithsonian pathology of an urban slave population from
Institution Press; p. 333–343. New Orleans. American Journal of Physical
Owsley DW, Bass WM. 1979. A demographic Anthropology 74:185–197.
analysis of skeletons from the Larson site Owsley DW, Rose JC, editors. 1997. Bioarcheology
(39WW2), Walworth County, South Dakota: of the North Central United States. Arkansas
Vital Statistics. American Journal of Physical Archeological Survey Research Series,
Anthropology 51:145–154. No. 49.
Owsley DW, Berryman HE, Bass WM. 1977. Owsley DW, Slutzky GD, Guagliardo MF, Deitrick
Demographic and Osteological Evidence for LM. 1981. Interpopulation relationships of four
Warfare at the Larson Site, South Dakota. post-contact Coalescent sites from South
Plains Anthropological Society Memoir, Dakota: Four Bear (39DW2), Oahe Village
No. 13: 119–131. (39HU2), Stony Point Village (39ST235) and
Owsley DW, Bruwelheide KS, Cartmell LW Sret al. Swan Creek (39WW7). In: Jantz RL, Ubelaker
2006. The man in the coffin: an DH, editors. Progress in Skeletal Biology of
interdisciplinary effort to name the past. Plains Populations. Plains Anthropologist
Historical Archaeology 40:89–108. Memoir, No.17; p. 31–42.
Owsley DW, Gill GW, Owsley SD. 1994. Biological Owsley DW, Symes SA. 1981. Morphological
effects of European contact on Easter Island. differences between Mandan and historic
In: Larsen CS, Milner GR, editors. In the Wake Arikara crania. In: Jantz RL, Ubelaker DH,
of Contact: Biological Responses to Conquest. editors. Progress in Skeletal Biology of Plains
New York, NY: Wiley-Liss; p. 161–177. Populations. Plains Anthropologist Memoir,
Owsley DW, Jantz RL. 1978. Intracemetery No.17; p. 49–56.
morphological variation in Arikara crania from Oxenham M. 2006. Biological responses to change
the Sully site (39SL4), Sully County, South in prehistoric Viet Nam. Asian Perspectives
Dakota. Plains Anthropologist 23:139–147. 45:212–239.
Owsley DW, Jantz RL. 1985. Long bone lengths Oxenham M, Cuong NL, Thuy NK. 2006. The oral
and gestational age distributions of post- health consequences of the adoption and
contact Arikara Indian perinatal infant intensification of agriculture in Southeast Asia.
skeletons. American Journal of Physical In: Oxenham MF, Tayles N, editors.
Anthropology 68:321–328. Bioarchaeology of Southeast Asia. Cambridge,
Owsley DW, Jantz RL, editors. 1994. Skeletal UK: Cambridge University Press; p. 263–289.
Biology in the Great Plains: Migration, Oxenham M, Tayles N, editors. 2006.
Warfare, Health, and Subsistence. Washington, Bioarchaeology of Southeast Asia. Cambridge,
DC: Smithsonian Institution Press. UK: Cambridge University Press.
Owsley DW, Mann RW, Murphy SP. 1991. Injuries, Oxenham MF, Dommett KM. 2011. Palaeohealth
surgical care, and disease. In: Pfeiffer S, at Man Bac. In: Oxenham MF, Matsumura H,
Williamson RF, editors. Snake Hill: An Dung DK, editors. Man Bac: The Excavation of
Investigation of a Military Cemetery from the a Late Neolithic Site in Northern Vietnam.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
536 References

Volume II: Biological Research. Canberra, Past and Present. Budapest/Szeged, Hungary:
Australia: ANU E-Press; p. 107–121. Golden Book Publishers and Tuberculosis
Oxenham MF, Thuy NK, Cuong NL. 2005. Skeletal Foundation; p. 533–539.
evidence for the emergence of infectious Pálfi G, Zink A, Haas C, et al. 2002. Historical and
disease in Bronze and Iron Age northern palaeopathological evidence of leprosy in
Vietnam. American Journal of Physical Hungary. In: Roberts CA, Lewis ME,
Anthropology 126:359–376. Manchester K, editors. The Past and Present of
Özer BK, Sağir M, Özer I. 2011. Secular changes in Leprosy: Archaeological, Historical,
the height of the inhabitants of Anatolia Palaeopathological and Clinical Approaches.
(Turkey) from the 10th millennium B.C. to the Oxford, UK: British Archaeological Reports,
20th century A.D. Economics and Human International Series, p. 205–212.
Biology 9:211–219. Palkovich AM. 1980. The Arroyo Hondo Skeletal
Oziegbe EO, Esan TA, Oyedele TA. 2014. and Mortuary Remains. Sante Fe, NM: School
Emergence chronology of permanent teeth in of American Research Press.
Nigerian children. American Journal of Palkovich AM. 1981. Tuberculosis epidemiology
Physical Anthropology 153:506–511. in two Arikara skeletal samples: a study of
Pääbo S. 1985. Molecular cloning of ancient disease impact. In: Buikstra JE, editor.
Egyptian mummy DNA. Nature 314:644–645. Prehistoric Tuberculosis in the Americas.
Pääbo S, Poinar H, Serre D, et al. 2004. Genetic Center for American Archeology, Scientific
analyses from ancient DNA. Annual Review of Papers, No.5; p. 161–175.
Genetics 38:645–679. Palkovich AM. 1983. A comment on Mobley’s
Paine RR. 2000. If a population crashes in “Demographic Structure of Pecos Indians.”
prehistory, and there is no paleodemographer American Antiquity 48:142–149.
there to hear it, does it make a sound? Palkovich AM. 1987. Endemic disease patterns in
American Journal of Physical Anthropology paleopathology: porotic hyperostosis. American
112:181–190. Journal of Physical Anthropology 74:527–537.
Pálfi G. 1991. The first osteoarchaeological Palkovich AM. 2008. Rickets, community
evidence of leprosy in Hungary. International dynamics, and gender relations at Arroyo
Journal of Osteoarchaeology 1:99–102. Hondo, a fourteenth century ancestral pueblo:
Pálfi G. 1992. Traces des activités sur les squelettes reanalysis and new insights. In: Stodder ALW,
des anciens Hongrois. Bulletins et Mémoires editor. Reanalysis and Reinterpretation in
de la Société‚ d’Anthropologie de Paris Southwestern Bioarchaeology. Arizona State
4:209–231. University Anthropological Research Papers,
Pálfi G, Bérato J, Dutour O. 1994. No. 59; p. 144–149.
Paleopathological data of the osteological Palmer AR, Strobeck C. 2003. Fluctuating
series from Costebelle, Hyeres (3rd–6th century asymmetry analysis revisited. In: Polak M,
A.D.). In: Dutour O, Pálfi G, Bérato JP, editors. editor. Developmental Instability: Causes and
L’Origin de la Syphilis en Europe: Avant ou Consequences. Oxford, UK: Oxford University
Après 1493? Toulon, France: Centre Press; p. 279–319.
Archeologique du Var, Éditions Errance, Panush RS, Brown DG. 1987. Exercise and
p. 125–132. arthritis. Sports Medicine 4:54–64.
Pálfi G, Dutour O, Deak J, Hutas I. 1999. Papageorgopoulou C, Suter SK, Rühli FJ,
Tuberculosis: Past and Present. Budapest, Siegmund F. 2011. Harris lines revisited:
Hungary: Golden Book Publishers and prevalence, comorbidities, and possible
Tuberculosis Foundation. etiologies. American Journal of Human Biology
Pálfi G, Marcsik A. 1999. Paleoepidemiological 23:381–391.
data of tuberculosis in Hungary. In: Pálfi P, Papathanasiou A. 2001. A Bioarchaeological
Dutour O, Deak J, Hutas I, editors. Tuberculosis: Analysis of Neolithic Alepotrypa Cave, Greece.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 537

British Archaeological Reports, International nineteenth-century free Black population in


Series, No. 961. Philadelphia, Pennsylvania. In: Buikstra JE,
Papathanasiou A. 2003. Stable isotope analysis in editor. A Life in Science: Papers in Honor of
Neolithic Greece and possible implications on J. Lawrence Angel. Center for American
human health. International Journal of Archeology, Scientific Papers, No. 6;
Osteoarchaeology 13:314–324. p. 138–170.
Papathanasiou A. 2005. Health status of the Paschetta C, de Azevedo S, Castillo L, et al. 2010.
Neolithic population of Alepotrypa Cave, The influence of masticatory loading on
Greece. American Journal of Physical craniofacial morphology; a test case across
Anthropology 126:377–390. technological transitions in the Ohio Valley.
Papathanasiou A. 2011. Health, diet and social American Journal of Physical Anthropology
implications in Neolithic Greece from the study 141:297–314.
of human osteological material. In: Pinhasi R, Pastor R. 1992. Dietary adaptations and dental
Stock J, editors. Human Bioarchaeology of the microwear in Mesolithic and Chalcolithic
Transition to Agriculture. Chichester, UK: South Asia. In: Lukacs JR, editor. Culture,
Wiley-Blackwell; p. 87–106. Ecology and Dental Anthropology. Journal of
Papathanasiou A, Larsen CS, Norr L. 2000. Human Ecology 2(Special Issue); p. 215–228.
Bioarchaeological inferences from a Neolithic Pastor R. 1993. Dental Microwear among
ossuary from Alepotrypa Cave, Diros, Greece. Prehistoric Inhabitants of the Indian
International Journal of Osteoarchaeology Subcontinent: A Quantitative and Comparative
10:210–228. Analysis. PhD Dissertation, University of
Parham KR, Scott GT. 1980. Porotic hyperostosis: Oregon, Eugene, OR.
a study of disease and culture at Toqua Pate FD. 1994. Bone chemistry and paleodiet.
(40MR6), a late Mississippian site in eastern Journal of Archaeological Method and Theory
Tennessee. In: Willey P, Smith FH, editors. The 1:161–209.
Skeletal Biology of Aboriginal Populations in Pate D, Hutton JT. 1988. The use of soil chemistry
the Southeastern United States. Tennessee data to address post-mortem diagenesis in
Anthropological Association, Miscellaneous bone mineral. Journal of Archaeological
Paper, No. 5; p. 39–51. Science 15:729–739.
Park EA, Richter CP. 1953. Transverse lines in Pate FD, Hutton JT, Norrish K. 1989. Ionic
bone: the mechanism of their development. exchange between soil solution and bone:
Bulletin of the Johns Hopkins Hospital toward a predictive model. Applied
93:364–388. Geochemistry 4:303–316.
Parker S, Roberts C, Manchester K. 1985–1986. Pate FD, Hutton RA, Gould RA, Pretty GL. 1991.
A review of British trepanations with reports Alterations of in vivo elemental dietary
on two new cases. Ossa 12:141–157. signatures in archaeological bone: evidence
Parker Pearson, M. 2000. The Archaeology of from the Roonka Flat Dune, South Australia.
Death and Burial. College Station, TX: A&M Archaeology of Oceania 26:58–69.
University Press. Patterson DK Jr. 1984. A Diachronic Study of
Parks DR. 1979. The Northern Caddoan languages: Dental Palaeopathology and Attritional Status
their subgrouping and time depths. Nebraska of Prehistoric Ontario Pre-Iroquois and
History 60:197–213. Iroquois Populations. Archaeological Survey of
Parr RL, Carlyle SW, O’Rourke DH. 1996. Ancient Canada, Mercury Series Paper, No. 122.
DNA analysis of Fremont Amerindians of the Patz JA, Hulme M, Rosenzweig C, et al. 2002.
Great Salt Lake wetlands. American Journal of Regional warming and malaria resurgence.
Physical Anthropology 99:507–518. Nature 99:12506–12508.
Parrington M, Roberts DG. 1990. Demographic, Pauwels F. 1976. Biomechanics of the Normal and
cultural, and bioanthropological aspects of a Diseased Hip. Berlin, Germany: Springer-Verlag.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
538 References

Paynter KJ, Grainger RM. 1956. The relation of Crane-Kramer GMM, editors. Ancient Health:
nutrition to the morphology and size of rat Skeletal Indicators of Agricultural and
molar teeth. Journal of the Canadian Dental Economic Intensification. Gainesville, FL:
Association 22:519–531. University Press of Florida; p. 92–112.
Pearson JA. 2013. Human and animal diet as Pechenkina EA, Xiaolin X, Wenquan F. 2013.
evidenced by stable carbon and nitrogen Trajectories of health in early farming
isotope analysis. In: Hodder I, editor. Humans communities of East Asia. In: Pechenkina K,
and Landscapes of Çatalhöyük. Los Angeles, Oxenham M, editors. Bioarchaeology of East
CA: Cotsen Institute of Archaeology; Asia: Movement, Contact, Health. Gainesville,
p. 271–299. FL: University Press of Florida; p. 444–481.
Pearson JA, Grove M, Özbek M, Hongo H. 2013. Pechenkina K, Oxenham M, editors. 2013.
Food and social complexity at Çayönü Tepesi, Bioarchaeology of East Asia: Movement,
southeastern Anatolia: stable isotope evidence Contact, Health. Gainesville, FL: University
of differentiation in diet according to burial Press of Florida.
practice and sex in the early Neolithic. Journal Peck JJ. 2013. Status, health, and lifestyle in
of Anthropological Archaeology 32:180–189. Middle Iron Age Britain: a bioarchaeological
Pearson JA, Hedges REM, Molleson TI, Özbek M. study of elites and non-elites from East
2010. Exploring the relationship between Yorkshire, Northern England. International
weaning and infant mortality: an isotope case Journal of Paleopathology 3:83–94.
study from Aşıklı Höyük and Çayönü Tepesi. Peck JJ, Stout SD. 2009. The effects of total hip
American Journal of Physical Anthropology arthroplasty on the structural and
143:448–457. biomechanical properties of adult bone.
Pearson OM, Buikstra JE. 2006. Behavior and the American Journal of Physical Anthropology
bones. In: Buikstra JE, Beck LA, editors. 138:221–230.
Bioarchaeology: The Contextual Analysis of Pedersen PO. 1944. Dental notes and a chapter on
Human Remains. Amsterdam, the Netherlands: the dentition. In: Broste K, Fischer-Moller K,
Elsevier; p. 207–225. editors. The Mediaeval Norsemen at Gardar:
Pechenkina EA, Ambrose SH, Xiaolin M, Benfer Anthropological Investigations. Meddelelser
RA Jr. 2005. Reconstructing northern Chinese om Grønland, Man & Society, 89; p. 1–62.
Neolithic subsistence practices by isotopic Pedersen PO. 1947. Dental investigations of
analysis. Journal of Archaeological Science Greenland Eskimos. Proceedings of the Royal
32:1176–1189. Society of Medicine 40:726–732.
Pechenkina EA, Benfer RA Jr, Ma X. 2007. Diet Pedersen PO. 1952. Some dental aspects of
and health in the Neolithic of the Wei and anthropology. Dental Record 72:170–178.
Middle Yellow River basins, northern China. In: Pedersen PO, Jakobsen J. 1989. Teeth and jaws of
Cohen MN, Crane-Kramer GMM, editors. the Qilakitsoq mummies. In: Hart Hansen JP,
Ancient Health: Skeletal Indicators of Gulløv HC, editors. The Mummies from
Agricultural and Economic Intensification. Qilakitsoq – Eskimos in the 15th Century.
Gainesville, FL: University Press of Florida; Meddelelser om Grønland, Man & Society, 12;
p. 255–272. p. 112–130.
Pechenkina EA, Benfer RAJ, Zhijun W. 2002. Diet Pedersen PO, Scott DB. 1951. Replica studies of the
and health change with the intensification of surfaces of teeth from Alaskan Eskimo, West
millet agriculture at the end of Chinese Greenland natives, and American Whites. Acta
Neolithic. American Journal of Physical Odontologica Scandinavica 9:262–292.
Anthropology 117:15–36. Peregrine PN. 2001. Matrilocality, corporate
Pechenkina EA, Vradenburg JA, Benfer RA Jr, strategy, and the organization of production in
Farnum JF. 2007. Skeletal biology of the the Chacoan world. American Antiquity
central Peruvian coast. In: Cohen MN, 66:36–46.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 539

Perez SI. Bernal V, Gonzalez PN, Sardi M, Politis Peterson R, Vucetich JA, Fenton G, Drummer TD,
GG. 2009. Discrepancy between cranial and Larsen CS. 2010. Ecology of arthritis. Ecology
DNA data of early Americans: implications for Letters 13:1124–1128.
American peopling. PLoS One 4(5)e5746:1–11. Pezo Lanfranco L, Eggers S. 2010. The usefulness
Perry EM. 2007. Is bioarchaeology a handmaiden of caries frequency, depth, and location in
to history? Developing a historical determining cariogenecity and past
bioarchaeology. Journal of Anthropological subsistence: a test on early and later
Archaeology 26:486–515. agriculturalists from the Peruvian coast.
Perry EM. 2008. Gender, labor, and inequality at American Journal of Physical Anthropology
Grasshopper Pueblo. In: Stodder ALW, editor. 143:75–91.
Reanalysis and Reinterpretation in Pfeiffer S. 1980. Age changes in the external
Southwestern Bioarchaeology. Arizona State dimensions of adult bone. American Journal of
University Anthropological Research Papers, Physical Anthropology 52:529–532.
No. 59; p. 151–166. Pfeiffer S. 1984. Paleopathology in an Iroquoian
Perry MA, Coleman DS, Delhopital N. 2008. ossuary, with special reference to tuberculosis.
Mobility and exile at 2nd century A.D. Khirbet American Journal of Physical Anthropology
edh-Dharih: strontium isotope analysis of 65:181–189.
human migration in western Jordan. Pfeiffer S. 1991. Rib lesions and New World
Geoarchaeology 23:528–549. tuberculosis. International Journal of
Perry MA, Coleman DS, Dettman DL, al-Shiyab Osteoarchaeology 1:191–198.
AH. 2009. An isotopic perspective on the Pfeiffer S, Dudar JC, Austin S. 1989. Prospect Hill:
transport of Byzantine mining camp laborers skeletal remains from a 19th-century
into southwestern Jordan. American Journal of Methodist cemetery. Northeast Historical
Physical Anthropology 140:429–441. Archaeology 18:29–48.
Perry MA, Coleman DS, Dettman DL, al-Shiyab Pfeiffer S, Fairgrieve SI. 1994. Evidence from
AH. 2012. Condemned to metallum? ossuaries: the effect of contact on the health of
Illuminating life at the Byzantine mining camp Iroquoians. In: Larsen CS, Milner GR, editors.
at Phaeno in Jordan. In: Perry MA, editor. In the Wake of Contact: Biological Responses to
Bioarchaeology and Behavior: The People of the Conquest. New York, NY: Wiley-Liss; p. 47–61.
Ancient Near East. Gainesville, FL: University Pfeiffer SK, King P. 1983. Cortical bone formation
Press of Florida; p. 115–137. and diet among protohistoric Iroquoians.
Perzigian AJ. 1977. Fluctuating dental American Journal of Physical Anthropology
asymmetry: variation among skeletal 60:23–28.
populations. American Journal of Physical Pfeiffer SK, Lazenby RA. 1994. Low bone mass in
Anthropology 47:81–88. past and present aboriginal populations.
Perzigian AJ, Tench PA, Braun DJ. 1984. Advances in Nutritional Research 9:35–51.
Prehistoric health in the Ohio River valley. In: Phillips SM. 2003. Worked to the bone: the
Cohen MN, Armelagos GJ, editors. biomechnical consequences of “labor therapy”
Paleopathology at the Origins of Agriculture. at a nineteenth century asylum. In: Herring
Orlando, FL: Academic Press; p. 347–366. DA, Swedlund AC, editors. Human Biologists in
Peterson J. 1998. The Natufian hunting the Archives. Cambridge, UK: Cambridge
conundrum: spears, atlatls, or bows? University Press; p. 96–129.
Musculoskeletal and armature evidence. Philpott R. 1991. Burial Practices in Roman
International Journal of Osteoarchaeology Britain. British Archaeological Reports,
8:378–389. International Series, No. 219.
Peterson J. 2002. Sexual Revolutions: Gender and Pickering RB. 1984. Patterns of Degenerative Joint
Labor at the Dawn of Agriculture. Walnut Disease in Middle Woodland, Late Woodland,
Creek, CA: Altamira. and Mississippian Skeletal Series from the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
540 References

Lower Illinois Valley. PhD Dissertation, Skeleton, Second Edition. Chichester, UK:
Northwestern University, Evanston, IL. Wiley-Blackwell; p. 487–532.
Pierce LKC. 1987. A Comparison of the Pattern of Pietrusewsky M. 2010. A multivariate analysis of
Involvement of Degenerative Joint Disease measurements recorded in early and more
between an Agricultural and Non-Agricultural modern crania from East Asia and Southeast
Skeletal Series. PhD Dissertation, University of Asia. Quaternary International 211:42–54.
Tennessee, Knoxville, TN. Pietrusewsky M. 2013. Biological connections
Pietrusewsky M. 1988. Prehistoric Human across the Sea of Japan: a multivariate
Remains from Non Pa Kluay, Northeast comparison of ancient and more modern crania
Thailand. University of Otago Studies in from Japan, China, Korea, and Southeast Asia.
Prehistoric Anthropology, No. 17. In: Pechenkina E, Oxenham MF, editors.
Pietrusewsky M. 1989. A study of skeletal and Bioarchaeology of East Asia: Movement,
dental remains from Watom Island and Contact, Health. Gainesville, FL: University
comparisons with other Lapita people. Records Press of Florida; p. 144–178.
of the Australian Museum 41:235–292. Pietrusewsky M, Douglas MT. 1994. An
Pietrusewsky M. 1994. Pacific-Asian osteological assessment of health and disease
relationships: a physical anthropological in precontact and historic (1778) Hawai’i. In:
perspective. Oceanic Linguistics 33:407–429. Larsen CS, Milner GR, editors. In the Wake of
Pietrusewsky M. 1999. A multivariate Contact: Biological Responses to Conquest.
craniometric study of the inhabitants of the New York, NY: Wiley-Liss; p. 179–196.
Ryukyu islands and comparisons with cranial Pietrusewsky M, Douglas MT. 2002a.
series from Japan, Asia, and the Pacific. Intensification of agriculture at Ban Chiang: is
Anthropological Science 107:255–281. there evidence from the skeletons? Asian
Pietrusewsky M. 2000. Metric analysis of skeletal Perspectives 40:157–178.
remains: methods and applications. In: Pietrusewsky M, Douglas MT. 2002b. Ban Chiang:
Katzenberg MA, Saunders SR, editors. A Prehistoric Village Site in Northeast
Biological Anthropology of the Human Thailand. I. The Human Skeletal Remains.
Skeleton. New York, NY: Wiley-Liss; Philadelphia, PA: University of Pennsylvania.
p. 375–415. Pietrusewsky M, Ikehara-Quebral R. 2006. The
Pietrusewsky M. 2005. The physical anthropology bioarchaeology of the Vat Komnou cemetery,
of the Pacific, East Asia, and Southeast Asia: a Angkor Borei, Cambodia. Bulletin of the Indo-
multivariate craniometric analysis. In: Sagat L, Prehistory Association 26:86–97.
Blench R, Sanchez-Mazas A, editors. The Pihlstrom BL. 2001. Periodontal risk assessment,
Peopling of East Asia: Putting Together diagnosis and treatment planning.
Archaeology, Linguistics, and Genetics. Periodontology 25:37–58.
London, UK: Routledge Curzon; p. 201–229. Pihlstrom BL, Michalowicz BS, Johnson NW.
Pietrusewsky M. 2006. A multivariate 2005. Periodontal diseases. The Lancet
craniometric study of the prehistoric and 366:1809–1820.
modern inhabitants of Southeast Asia, East Pijoan CM, Mansilla J. 1990. Praticas rituales en el
Asia and surrounding regions: a human norte de Mesoamérica: evidencias en Electra,
kaleidoscope? In: Oxenham M, Tayles N, Villa de Reyes, San Luis Potosí. Arqueologia
editors. Bioarchaeology of Southeast Asia. 4:87–96.
Cambridge, UK: Cambridge University Press; Pijoan CM, Mansilla LJ. 1997. Evidence for human
p. 59–90. sacrifice, bone modification and cannibalism
Pietrusewsky M. 2008. Metric analysis of skeletal in ancient Mexico. In: Martin D, Frayer D,
remains: methods and applications. In: editors. Troubled Times: Violence and Warfare
Katzenberg MA, Saunders SR, editors. in the Past. Langhorne, PA: Gordon and
Biological Anthropology of the Human Breach; p. 217–239.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 541

Pilloud MA, Canzonieri C. 2012. The occurrence Pinhasi R, Timpson A, Thomas M, Šlaus M. 2014.
and possible etiology of spondylolysis in a pre- Bone growth, limb proportions and non-
contact California population. International specific stress in archaeological populations
Journal of Osteoarchaeology DOI:10.1002/ from Croatia. Annals of Human Biology
oa.2245. 41:127–137.
Pilloud MA, Jones TL, Schwitalla AW. 2014. The Pinhasi R, Von Cramon-Taubedel N. 2009.
bioarchaeological record of craniofacial Craniometric data supports demic diffusion
trauma in central California. In: Allen MW, model for the spread of agriculture into Europe.
Jones TL, editors. Re-examining a Pacified PLoS One 4(8):e6747.
Past: Violence and Warfare among Hunter- Piziali RL, Hight TK, Nagel DA. 1976. An extended
Gatherers. Walnut Creek, CA: Left Coast Press; structural analysis of long bones–application
p. 257–272. to the human tibia. Journal of Biomechanics
Pilloud MA, Larsen CS. 2011. “Official” and 9:695–701.
“practical” kin: inferring social and community Platt BS, Stewart RJC. 1962 Transverse trabeculae
structure from dental phenotypes at Neolithic and osteoporosis in bones in experimental
Çatalhöyük, Turkey. American Journal of protein-calorie deficiency. British Journal of
Physical Anthropology 145:519–530. Nutrition 16:483–495.
Pilot T. 1998. The periodontal disease problem: a Polak M. 2003. Developmental Instability: Causes
comparison between industrialized and and Consequences. New York, NY: Oxford
developing countries. International Dental University Press.
Journal 48(Suppl. 1):221–232. Polet C, Katzenberg MA. 2003. Reconstruction of
Pindborg JJ. 1982. Aetiology of developmental the diet in a mediaeval monastic community
defects not related to fluorosis. International from the coast of Belgium. Journal of
Dental Journal 32:123–134. Archaeological Science 30:525–533.
Pinhasi R, Bourbou C. 2008. How representative Polet C, Orban R, Herbosch A. 2000. Indicateurs de
are human skeletal assemblages for population stress et teneurs en éléments traces: exemple de
analysis? In: Pinhasi R, Mays S, editors. deux populations Mediévales de Belgique.
Advances in Human Paleopathology. New Bulletins et Mémoires de la Société
York, NY: John Wiley & Sons, p. 31–44. d’Anthropologie de Paris 12:247–278.
Pinhasi R, Eshed V, Shaw P. 2008. Evolutionary Polk JD, Demes B, Jungers WL, et al. 2000.
changes in the masticatory complex following A comparison of primate, carnivore, and
the transition to farming in the southern rodent limb bone cross-sectional properties:
Levant. American Journal of Physical are primates really unique? Journal of Human
Anthropology 135:136–148. Evolution 39:297–325.
Pinhasi R, Fort J, Ammerman AJ. 2005. Tracing Pollack D, Henderson AG. 2000. Insights into Fort
the origin and spread of agriculture in Europe. Ancient culture change: a view from south of
PLoS Biology 3:2220–2228. the Ohio River. In: Genheimer RA, editor.
Pinhasi R, Meiklejohn C. 2011. Dental reduction Cultures Before Contact: The Late Prehistory of
and the transition to agriculture in Europe. In: Ohio and Surrounding Regions. Columbus, OH:
Pinhasi R, Stock JT, editors. Human Ohio Archaeological Council; p. 194–215.
Bioarchaeology of the Transition to Agriculture. Pollard AM, Ditchfield P, McCullagh JSO, et al.
Chichester, UK: Wiley-Blackwell; p. 451–474. 2011. “These boots are made for walking”: the
Pinhasi R, Stock J, editors. 2011. The isotopic analysis of a C4 Roman inhumation
Bioarchaeology of the Agricultural Transition. from Gravesend, Kent, UK. American Journal
London, UK: Wiley-Blackwell. of Physical Anthropology 146:446–456.
Pinhasi R, Thomas MG, Hofreiter M, Currat M, Pollitzer WS, Anderson JJB. 1989. Ethnic and
Burger J. 2012. The genetic history of genetic differences in bone mass: a review with
Europeans. Trends in Genetics 28:496–505. a hereditary vs. environmental perspective.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
542 References

American Journal of Clinical Nutrition Powell ML. 1985. The analysis of dental wear and
50:1244–1259. caries for dietary reconstruction. In: Gilbert RI,
Popkin BM, Lasky T, Litvin J, Spicer D, Yamamoto Mielke JH, editor. The Analysis of Prehistoric
ME. 1986. The Infant-Feeding Triad: Infant, Diets. Orlando, FL: Academic Press; p. 307–338.
Mother and Household. New York, NY: Gordon Powell ML. 1986. Late prehistoric community
and Breach Science Publishers. health in the central deep south: biological and
Porro MA, Boano R, Spani F, Doro Geretto T. 1999. social dimensions of the Mississippian
Studi antropologici sulla popolazione. In: chiefdom at Moundville, Alabama. In: Levy JE,
Negro Ponzi Mancini MM, editor. San Michele editor. Skeletal Analysis in Southeastern
di Trino. Dal Villaggio Romano al Castello Archaeology. North Carolina Archaeological
Medievale, Volume 2. Council, No. 24; p. 127–149.
Poss R. 1984. Functional adaptation of the human Powell ML. 1988. Status and Health in Prehistory:
locomotor system to normal and abnormal A Case Study of the Moundville Chiefdom.
loading patterns. Calcified Tissue Research Washington, DC: Smithsonian Institution
36:151–161. Press.
Post DM. 2002. Using stable isotopes to estimate Powell ML. 1989. The people of Nodena. In: Morse
trophic position: models, methods, and DF, editor. Nodena: An Account of 90 Years of
assumptions. Ecology 83:703–718. Archeological Investigation in Southeast
Pott P. 1779. Remarks on the Kind of Palsy of the Mississippi County, Arkansas. Arkansas
Lower Limbs which is Frequently Found to Archeological Survey Research Series, No. 30;
Accompany a Curvature of the Spine. London, p. 65–95.
UK: J. Johnson. Powell ML. 1990. On the eve of conquest: life and
Potter RH, Nance WE, Pao-Lo Y, Davis WB. death at Irene Mound, Georgia. In: Larsen CS,
1976. A twin study of dental dimension. II. editor. The Archaeology of Santa Catalina De
Independent genetic determinants. Guale: 2. Biocultural Interpretations of a
American Journal of Physical Anthropology Population in Transition. Anthropological
44:397–412. Papers of the American Museum of Natural
Potter RHY, Rice JP, Dahlberg AA, Dahlberg T. History, No. 68; p. 26–35.
1983. Dental size traits within families: path Powell ML. 1991a. Endemic treponematosis and
analysis for first molar and lateral incisor. tuberculosis in the prehistoric southeastern
American Journal of Physical Anthropology United States: biological costs of chronic
61:283–289. endemic disease. In: Ortner DJ, Aufderheide
Powell JF. 2005. The First Americans: Race, AC, editor. Human Paleopathology: Current
Evolution, and the Origin of Native Americans. Syntheses and Future Options. Washington,
Cambridge, UK: Cambridge University Press. DC: Smithsonian Institution Press; p. 173–180.
Powell JF, Neves WA. 1999. Craniofacial Powell ML. 1991b. Ranked status and health in the
morphology of the First Americans: pattern Mississippian chiefdom at Moundville. In:
and process in the peopling of the New World. Powell ML, Bridges P, Mires AMW, editors.
Yearbook of Physical Anthropology What Mean These Bones? Studies in
42:153–188. Southeastern Bioarchaeology. Tuscaloosa, AL:
Powell JF, Steele DG. 1994. Diet and health of University of Alabama Press; p. 22–51.
Paleoindians: an examination of early Powell ML. 1992a. In the best of health? Disease
Holocene human dental remains. In: Sobolik and trauma among the Mississippian elite. In:
KD, editor. Paleonutrition: The Diet and Health Barker AW, Pauketat TR, editors. Lords of the
of Prehistoric Americans. Southern Illinois Southeast: Social Inequality and the Native
University at Carbondale, Center for Elites of Southeastern North America.
Archaeological Investigations, Occasional Archaeological Papers of the American
Paper, No. 22; p. 178–194. Anthropological Association, No. 3; p. 81–97.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 543

Powell ML. 1992b. Health and disease in the late Price B, Cameron N, Tobias PV. 1987. A further
prehistoric Southeast. In: Verano JW, Ubelaker search for a secular trend of adult body size in
DH, editor. Disease and Demography in the South African Blacks: evidence from the femur
Americas. Washington, DC: Smithsonian and tibia. Human Biology 59:467–475.
Institution Press; p. 41–53. Price D. 2008. Isotopes and human migration: case
Powell ML. 1994. Human skeletal remains from studies in biogeochemistry. In: Schutkowski H,
Ocmulgee National Monument. In: Hally DJ, editor. Between Biology and Culture.
editor. Ocmulgee Archaeology, 1936–1986. Cambridge, UK: Cambridge University Press;
Athens, GA: University of Georgia Press; p. 243–272.
p. 116–129. Price TD. 1989. Multi-element studies of
Powell ML. 2000. Ancient diseases, modern diagenesis in prehistoric bone. In: Price TD,
perspectives: treponematosis and tuberculosis editor. The Chemistry of Prehistoric Human
in the age of agriculture. In: Lambert PM, Bone. Cambridge, UK: Cambridge University
editor. Bioarchaeological Studies of Life in the Press; p. 17–39.
Age of Agriculture: A View from the Southeast. Price TD, Blitz J, Burton JH, Ezzo JA. 1992.
Tuscaloosa, AL: University of Alabama; Diagenesis in prehistoric bone: problems and
p. 6–34. solutions. Journal of Archaeological Science
Powell ML, Cook DC, editors. 2005a. The Myth 19:413–529.
of Syphilis: The Natural History of Price TD, Burton JH, Bentley RA. 2002. The
Treponematosis in North America. characterization of biologically available
Gainesville, FL: University Press of Florida. strontium isotope ratios for the study of
Powell ML, Cook DC. 2005b. Treponematosis: prehistoric migration. Archaeometry
inquiries into the nature of a protean disease. 44:117–135.
In: Powell ML, Cook DC, editors. The Myth Price TD, Burton JH, Wright LE, White CD,
of Syphilis: The Natural History of Longstaffe F. 2007. Victims of sacrifice:
Treponematosis in North America. Gainesville, isotopic evidence for place of origin. In: Tiesler
FL: University Press of Florida; p. 9–62. V, Cucina A, editors. New Perspectives on
Powell ML, Jacobi K, Danforth ME, Eisenberg LE. Human Sacrifice and Ritual Body Treatments
2005. Syphilis in Mound Builders’ bones. In: in Ancient Maya Society. New York, NY:
Powell ML, Cook DC, editors. The Myth of Springer; p. 263–292.
Syphilis: The Natural History of Price TD, Frei KM, Dobat AS, Lynnerup N, Bennike
Treponematosis in North America. Gainesville, P. 2011. Who was in Harold Bluetooth’s army?
FL: University Press of Florida; p. 117–161. Strontium isotope investigation of the
Powell ML, Rogers JD. 1980. Bioarchaeology of the cemetery at the Viking age fortress at
McCutchan-McLaughin Site. Oklahoma Trelleborg, Denmark. Antiquity 85:476–489.
Archaeological Survey, Studies in Oklahoma’s Price TD, Grupe G, Schroter P. 1994.
Past, No. 5. Reconstruction of migration patterns in the
Pratt LW. 1943. Experimental masseterectomy in Bell Beaker period by stable strontium isotope
the laboratory rat. Journal of Mammalogy analysis. Applied Geochemistry 9:413–417.
24:204–211. Price TD, Johnson CM, Ezzo JA, Ericson J, Burton
Pretty GL, Kricun ME. 1989. Prehistoric health JH. 1994. Residential mobility in the
status of the Roonka population. World prehistoric southwest United States: a
Archaeology 21:198–224. preliminary study using strontium isotope
Preus HR, Marvik OJ, Selvig KA, Bennike P. 2011. analysis. Journal of Archaeological Science
Ancient bacterial DNA (aDNA) in dental 21:315–330.
calculus from archaeological human remains. Price TD, Manzanilla L, Middleton WD. 2000.
Journal of Archaeological Science Immigration and the ancient city of
38:1827–1831. Teotihuacan in Mexico: a study using

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
544 References

strontium isotope ratios in human bone and evidence for age-related variation in diet from
teeth. Journal of Archaeological Science Isola Sacra, Italy. American Journal of Physical
27:903–913. Anthropology 128:2–13.
Price TD, Schoeninger MJ, Armelagos GJ. 1985. Pryor F. 1976. A Neolithic multiple burial from
Bone chemistry and past behavior: an overview. Fengate, Peterborough. Antiquity 50:232–233.
Journal of Human Evolution 14:419–447. Puech PF, Serratrice C, Leek FF. 1983. Tooth wear
Price TD, Tiesler V, Burton JH. 2006. Early African as observed in ancient Egyptian skulls. Journal
diaspora in colonial Campeche, Mexico: of Human Evolution 12:617–629.
strontium isotopic evidence. American Journal Raadsheer MC, van Eijden TM, van Ginkel FC,
of Physical Anthropology 130:485–590. Prahl-Andersen B. 1999. Contribution of jaw
Price WA. 1936. Eskimo and Indian field studies in muscle size and craniofacial morphology to
Alaska and Canada. Journal of the American human bite force magnitude. Journal of Dental
Dental Association 23:417–437. Research 78:31–42.
Priestley HI. 1937. A Historical, Political, and Radin EL. 1982. Mechanical factors in the causation
Natural Description of California by Pedro of osteoarthritis. Rheumatology 7:46–52.
Fages, Soldier of Spain. Ramona, Spain: Radin EL. 1983. The relationship between
Ballena Press. biological and mechanical factors in the
Prince RL, Price RI, Ho S. 1988. Forearm bone loss etiology of osteoarthritis. Journal of
in hemiplegia: a model for the study of Rheumatology 10(S9):20–21.
immobilization osteoporosis. Journal of Bone Radin EL, Burr DB, Caterson B, et al. 1991.
and Mineral Research 3:305–310. Mechanical determinants of osteoarthrosis.
Pritzker KPH. 2003. Pathology of osteoarthritis. Seminars in Arthritis and Rheumatism
In: Brandt KD, Doherty M, Lohmander LS, 21(3S2):12–21.
editors. Osteoarthritis, Second Edition. New Radin EL, Martin RB, Burr DB, et al. 1984. Effects
York, NY: Oxford University Press; p. 49–58. of mechanical loading on the tissues of the
Proffit WR, Fields HW, Sarver D. 2013. rabbit knee. Journal of Orthopaedic Research
Contemporary Orthodontics, Fifth Edition. 1:221–234.
St. Louis, MO: Elsevier Mosby. Radin EL, Orr RB, Kelman JL, Paul IL, Rose RM.
Prowse R, Schwarcz HP, Saunders S, Macchiarelli 1982. Effect of prolonged walking on concrete
R, Bondioli L. 2004. Isotopic paleodiet studies on the knees of sheep. Journal of Biomechanics
of skeletons from the Imperial-age cemetery of 15:487–492.
Isola Sacra. Journal of Archaeological Science Radin EL, Paul IL, Rose RM. 1972. Role of
31:256–271. mechanical factors in pathogenesis of primary
Prowse T, Schwarcz H, Garnsey P, et al. 2007. osteoarthritis. The Lancet 1:519–522.
Isotopic evidence for age-related immigration Radosevich SC. 1989. Diet or Diagenesis? An
to Imperial Rome. American Journal of Evaluation of the Trace Element Analysis of
Physical Anthropology 132:510–519. Bone. PhD Dissertation, University of Oregon,
Prowse TL. 2011. Diet and dental health through Eugene, OR.
the life course in Roman Italy. In: Agarwal SC, Radosevich SC. 1993. The six deadly sins of trace
Glencross BA, editors. Social Bioarchaeology. element analysis: a case of wishful thinking in
Chichester, UK: Wiley-Blackwell; p. 412–439. science. In: Sandford MK, editor. Investigations
Prowse TL, Saunders SR, Schwarcz HP, et al. 2008. of Ancient Human Tissue: Chemical Analyses
Isotopic and dental evidence for infant and in Anthropology. Langhorne, PA: Gordon and
young child feeding practices in an imperial Breach Science Publishers; p. 269–332.
Roman skeletal sample. American Journal of Raemsch CA. 1995. Craniometric Variation within
Physical Anthropology 137:294–308. Skeletal Samples of Diverse Biological
Prowse TL, Schwarcz HP, Saunders SR, Composition. PhD Dissertation, State
Macchiarelli R, Bondioli L. 2005. Isotopic University of New York, Albany, NY.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 545

Raff J, Cook DC, Kaestle F. 2006. Tuberculosis in In: Powell ML, Bridges PS, Mires AMW, editors.
the New World: a study of ribs from the Schild What Mean These Bones? Studies in
Mississippian population, west-central Illinois. Southeastern Bioarchaeology. Tuscaloosa, AL:
Memórias do Instituto Oswaldo Cruz, Rio de University of Alabama Press; p. 148–164.
Janeiro 101(Supplement II):25–27. Rathbun TA, Sexton J, Michie J. 1980. Disease
Raff J, Tackney JJ, O’Rourke DH. 2010. South from patterns in a formative period South Carolina
Alaska: a pilot aDNA study of genetic history population. Tennessee Anthropological
on the Alaska peninsula and eastern Aleutians. Association, No. 5; p.52–74.
Human Biology 82:677–694. Rathbun TA, Steckel RH. 2002. The health of
Raff JA. 2008. An Ancient DNA Perspective on the slaves and free blacks in the East. In: Steckel
Prehistory of the Lower Illinois Valley. PhD RH, Rose JC, editors. The Backbone of History:
Dissertation, Indiana University, Bloomington, Health and Nutrition in the Western
IN. Hemisphere. New York, NY: Cambridge
Raff JA, Bolnick DA, Tackney J, O’Rourke DH. University Press; p. 208–225.
2011. Ancient DNA perspectives on American Ravosa MJ, Lopez EK, Menegaz RA, et al. 2008.
colonization and history. American Journal of Adaptive plasticity in the mammalian
Physical Anthropology 146:503–514. masticatory complex: you are what, and how,
Rafferty KL, Ruff CB. 1994. Articular structure and you eat. In: Vinyard CJ, Ravosa MJ, Wall CE,
function in Hylobates, Colobus, and Papio. editors. Primate Craniofacial Function and
American Journal of Physical Anthropology Biology. New York, NY: Springer; p. 293–328.
94:395–408. Rea AM. 1986. Black vultures and human victims:
Rafi A, Spigelman M, Stanford J. 1994. DNA of archaeological evidence from Pacatnamu. In:
Mycobacterium leprae detected by PCR in Donnan CB, Cock GA, editors. The Pacatnamu
ancient bone. International Journal of Papers, Volume 1. Los Angeles, CA: Museum of
Osteoarchaeology 4:287–290. Cultural History, University of California;
Raoult D, Aboudharam G, Crubézy E, et al. 2000. p. 139–144.
Molecular identification by “suicide PCR” of Reber EA. 2006. A hard row to hoe: changing
Yersinia pestis as the agent of Medieval Black maize use in the American Bottom and
Death. Proceedings of the National Academy of surrounding areas. In: Staller J, Benz B, editors.
Sciences 97:12800–12803. Histories of Maize. Amsterdam, the
Rasmussen KL, Boldsen JL, Kristensen HK, et al. Netherlands: Elsevier; p. 236–248.
2008. Mercury levels in Danish Medieval Redford DB. 1992. Egypt, Canaan and Israel in
human bones. Journal of Archaeological Ancient Times. Princeton, NJ: Princeton
Science 35:2295–2306. University Press.
Rasmussen M, Anzick SL, Waters MR, et al. 2014. Redmond EM. 1994. Tribal and Chiefly Warfare in
The genome of a Late Pleistocene human from South America. University of Michigan,
a Clovis burial site in western Montana. Nature Memoirs of the Museum of Anthropology,
506:225–229. No. 28.
Rathbun TA. 1984. Skeletal pathology from the Reed DM. 1994. Ancient Maya diet at Copán,
Paleolithic through the Metal Ages in Iran and Honduras, as determined through the analysis
Iraq. In: Cohen MN, Armelagos GJ, editors. of stable carbon and nitrogen isotopes. In:
Paleopathology at the Origins of Agriculture. Sobolik KD, editor. Paleonutrition: The Diet
Orlando, FL: Academic Press; p. 137–167. and Health of Prehistoric Americans. Southern
Rathbun TA. 1987. Health and disease at a South Illinois University at Carbondale, Center for
Carolina plantation: 1840–1870. American Archaeological Investigation, Occasional
Journal of Physical Anthropology 74:239–253. Paper, No. 22; p. 210–221.
Rathbun TA, Scurry JD. 1991. Status and health in Reed DM. 1999. Cuisine from Hun-Nal-Ye. In:
colonial South Carolina: Belleview Plantation. White CD, editor. Reconstructing Ancient Maya

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
546 References

Diet. Salt Lake City, UT: University of Utah DH, editor. Columbian Consequences, Volume
Press; p. 183–196. 2: Archaeological and Historical Perspectives
Reich D, Patterson N, Campbell D, et al. 2012. on the Spanish Borderlands East. Washington,
Reconstructing Native American population DC: Smithsonian Institution Press; p. 543–554.
history. Nature 488:370–375. Relethford J. 2010. The study of human population
Reichs KJ. 1989. Treponematosis: a possible case genetics. In: Larsen CS, editor. A Companion to
from the late prehistoric of North Carolina. Biological Anthropology. Malden, MA:
American Journal of Physical Anthropology Wiley-Blackwell; p. 74–87.
79:289–303. Relethford JH. 1994. Craniometric variation
Reid DJ, Dean MC. 2000. The timing of linear among modern human populations.
hypoplasias on human anterior teeth. American Journal of Physical Anthropology
American Journal of Physical Anthropology 95:53–62.
113:135–140. Relethford JH. 2012. Human Population Genetics.
Reid DJ, Dean MC. 2006. Variation in modern Chichester, UK: Wiley-Blackwell.
human enamel formation times. Journal of Relethford JH, Mahoney MC. 1991. Relationship
Human Evolution 50:329–346. between population density and rates of injury
Reinhard KJ, Ghazi AM. 1992. Evaluation of lead mortality in New York State (exclusive of New
concentrations in Nebraska skeletons using York City). American Journal of Human
ICP-MS. American Journal of Physical Biology 3:111–118.
Anthropology 89:183–195. Resnick D, Greenway G. 1982. Distal femoral
Reinhard KJ, Tieszen L, Sandness KL, et al. 1994. cortical defects, irregularities, and excavations.
Trade, contact, and female health in northeast Radiology 143:345–354.
Nebraska. In: Larsen CS, Milner GJ, editors. In Retzius G. 1900. Crania Suecica Antiqua.
the Wake of Contact: Biological Responses to Stockholm, Sweden.
Conquest. New York, NY: Wiley-Liss; p. 63–74. Rewekant A. 2001. Do environmental disturbances
Reinhardt GA. 1983. Relationships between of an individual’s growth and development
attrition and lingual tilting in human teeth. influence later bone involution processes?
American Journal of Physical Anthropology A study of two medieval populations.
61:227–237. International Journal of Osteoarchaeology
Reitsema LJ. 2013. Beyond diet reconstruction: 11:433–443.
stable isotope application to human Rhodes JA, Knüsel CJ. 2005. Activity-related
physiology, health, and nutrition. American skeletal change in Medieval humeri: cross-
Journal of Human Biology 25:445–456. sectional and architectural alterations.
Reitsema LJ, Crews DE, Polcyn M. 2010. American Journal of Physical Anthropology
Preliminary evidence for Medieval Polish diet 128:536–546.
from carbon and nitrogen stable isotopes. Ribot I, Roberts C. 1996. A study of non-specific
Journal of Archaeological Science stress indicators and skeletal growth in two
37:1413–1423. Mediaeval subadult populations. Journal of
Reitsema LJ, Vercellotti G. 2012. Stable isotope Archaeological Science 23:67–79.
evidence for sex- and status-based variations Ricaut F-X, Auriol V, von Cramon-Taubadel N,
in diet and life history at Medieval Trino et al. 2010. Comparison between
Vercellese, Italy. American Journal of Physical morphological and genetic data to estimate
Anthropology 148:589–600. biological relationships: the case of the Egyin
Reitz EJ. 1988. Evidence for coastal adaptation in Gol necropolis (Mongolia). American Journal
Georgia and South Carolina. Archaeology of of Physical Anthropology 143:355–364.
Eastern North America 16:137–158. Ricaut F-X, Kolodesnikov S, Keyser-Tracqui C,
Reitz EJ. 1990. Zooarchaeological evidence for et al. 2006. Molecular genetic analysis of
subsistence at La Florida missions. In: Thomas 400-year-old human remains found in two

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 547

Yakut burial sites. American Journal of magnesium in archaeological remains.


Physical Anthropology 129:55–63. American Journal of Physical Anthropology
Richards LC. 1984. Principal axis analysis of Supplement 22:197.
dental attrition data from two Australian Richtsmeier JT, Cheverud JM, Lele S. 1992.
Aboriginal populations. American Journal of Advances in anthropological morphometrics.
Physical Anthropology 65:5–13. Annual Review of Anthropology
Richards LC. 1990. Tooth wear and 21:283–305.
temporomandibular joint change in Australian Richtsmeier JT, Cole TM, Lele SR. 2005. An
Aboriginal populations. American Journal of invariant approach to the study of fluctuating
Physical Anthropology 82:377–384. asymmetry: developmental instability in a
Richards LC, Brown T. 1981. Dental attrition and mouse model for Down syndrome. In: Slice DE,
degenerative arthritis of the editor. Modern Morphometrics in Physical
temporomandibular joint. Journal of Oral Anthropology. New York, NY: Kluwer
Rehabilitation 8:293–307. Academic/Plenum Publishers; p. 187–212.
Richards LC, Miller SLJ. 1991. Relationships Riesenfeld A. 1970. The effect of environmental
between age and dental attrition in Australian factors on tooth development: an experimental
Aborigines. American Journal of Physical investigation. Acta Anatomica 77:188–215.
Anthropology 84:159–164. Rife JL. 2012. Isthmia, IX: The Roman and
Richards M, Montgomery J. 2012. Isotope analysis Byzantine Graves and Human Remains.
and paleopathology: a short review and future Princeton, NJ: American School of Classical
developments. In: Buikstra JE, Roberts C, Studies at Athens.
editors. The Global History of Paleopathology: Riggs P. 1994. The standard of living in Scotland,
Pioneers and Prospects. New York, NY: Oxford 1800–1850. In: Komlos J, editor. Stature,
University Press; p. 718–731. Living Standards, and Economic Development:
Richards MP, Hedges REM, Molleson TI, Vogel JC. Essays in Anthropometric History. Chicago, IL:
1998. Stable isotope analysis reveals variations University of Chicago Press; p. 60–75.
in human diet in the Poundbury Camp Rightmire GP. 1970. Bushman, Hottentot and
cemetery site. Journal of Archaeological South African Negro crania studied by distance
Science 25:1247–1252. and discrimination. American Journal of
Richards MP, Mays S, Fuller BT. 2002. Stable Physical Anthropology 33:169–196.
carbon and nitrogen isotope values of bone Rightmire GP. 1976. Multidimensional scaling and
and teeth reflect weaning age at the Medieval the analysis of human biological diversity in
Wharram Percy site, Yorkshire, UK. American Subsaharan Africa. American Journal of
Journal of Physical Anthropology Physical Anthropology 44:445–452.
119:2005–2010. Rightmire GP. 1984. Human skeletal remains from
Richards MP, Price TD, Koch E. 2003. Mesolithic eastern Africa. In: Clark JD, Brandt SA, editors.
and Neolithic subsistence in Denmark: new From Hunters to Farmers: The Causes and
stable isotope data. Current Anthropology Consequences of Food Production in Africa.
44:288–295. Berkeley, CA: University of California Press;
Richards MP, Schulting RJ, Hedges REM. 2003. p. 191–199.
Sharp shift in diet at onset of Neolithic. Nature Rizk OT, Amugongo SK, Mahaney MC, Hlusko LJ.
425:366. 2008. The quantitative genetic analysis of
Richman EA, Ortner DJ, Schulter-Ellis FP. 1979. primate dental variation: history of the
Differences in intracortical remodeling in three approach and prospects for the future. In: Irish
aboriginal populations: possible dietary JD, Nelson GC, editors. Technique and
factors. Calcified Tissue Research 28:209–214. Application in Dental Anthropology.
Richtsmeier JS, Sheridan SG. 1996. Elemental Cambridge, UK: Cambridge University Press;
analysis of cribra orbitalia: iron and p. 317–346.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
548 References

Robb J. 1994. Gender contradictions, moral Anthropology. British Archaeological


coalitions and inequality in prehistoric Italy. Reports, International Series, 291;
Journal of European Archaeology 2:20–49. p. 137–147.
Robb J. 1997a. Intentional tooth removal in Roberts C. 1991. Trauma and treatment in the
Neolithic Italian women. Antiquity British Isles in the historic period: a design for
71:659–669. multidisciplinary research. In: Ortner DJ,
Robb J. 1997b. Violence and gender in early Italy. Aufderheide AC, editors. Human
In: Martin D, Frayer D, editors. Troubled Times: Paleopathology: Current Syntheses and Future
Violence and Warfare in the Past. Langhorne, Options. Washington, DC: Smithsonian
PA: Gordon and Breach; p. 111–144. Institution Press; p. 225–240.
Robb J, Bigazzi R, Lazzarini L, Scarsini C, Sonego Roberts C. 1993. Pre-Columbian syphilis in
F. 2001. Social “status” and biological “status”: England in a well-preserved adult female
a comparison of grave goods and skeletal skeleton from Gloucester. Journal of
indicators from Pontecagnano. American Palaeopathology 5:111.
Journal of Physical Anthropology Roberts C. 2012. Re-emerging infections:
115:213–222. developments in bioarchaeological
Robbins DM, Rosenberg KR, Ruff CB. 1989. contributions to understanding tuberculosis
Activity patterns in late Middle Woodland, today. In: Grauer AL, editor. A Companion to
Delaware. American Journal of Physical Paleopathology. Chichester, UK: Wiley-
Anthropology 78:290–291. Blackwell; p. 434–457.
Robbins G, Sciulli PW, Blatt SH. 2010. Estimating Roberts C, Cox M. 2003. Health & Disease in
body mass in subadult human skeletons. Britain: From Prehistory to the Present Day.
American Journal of Physical Anthropology Phoenix Mill, UK: Sutton Publishing.
143:146–150. Roberts C, Cox M. 2007. The impact of economic
Robbins LM. 1978. Yawslike disease processes in a intensification and social complexity on
Louisiana shell mound population. Medical human health in Britain from 6000 BP
College of Virginia Quarterly 14:24–31. (Neolithic) and the introduction of farming
Robbins Schug G. 2011a. Don’t throw out the baby to the mid-nineteenth century AD. In: Cohen
with the bathwater: estimating fertility from MN, Crane-Kramer GMM, editors. Ancient
subadult skeletons. International Journal of Health: Skeletal Indicators of Agricultural
Osteoarchaeology 21:717–721. and Economic Intensification. Gainesville,
Robbins Schug G. 2011b. Bioarchaeology and FL: University Press of Florida; p. 149–163.
Climate Change: A View from South Asian Roberts C, Ingham S. 2008. Using ancient DNA
Prehistory. Gainesville, FL: University Press of analysis in palaeopathology: a critical analysis
Florida. of published papers, with recommendations for
Robbins Schug G, Gray K, Mushrif-Tripathy V, future work. International Journal of
Sankhyan AR. 2012. A peaceful realm? Trauma Osteoarchaeology 18:600–613.
and social differentiation at Harappa. Roberts C, Lucy D, Manchester K. 1994.
International Journal of Paleopathology Inflammatory lesions of ribs: an analysis of the
2:136–147. Terry Collection. American Journal of Physical
Robbins Schug G, Tripathy VM, Misra VN, et al. Anthropology 95:169–182.
2009. Ancient skeletal evidence for leprosy in Roberts C, Manchester K. 2005. The Archaeology of
India (2000 B.C.). PLoS ONE 4:1–8. Disease, Third Edition. Phoenix Mill, UK:
Robert RA. 1947. Chronic Structural Low Sutton Publishing.
Backache due to Low-Back Structural Roberts, CA. 1999. Rib lesions and tuberculosis:
Derangements. London, UK: H.K. Lewis. the current state of play. In: Pálfi G, Dutour O,
Roberts C. 1986. Leprogenic odontodysplasia. In: Deak J, Hutas I, editors. Tuberculosis: Past and
Cruwys E, Foley RA, editors. Teeth and Present. Budapest/Szeged, Hungary: Golden

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 549

Book Publishers and Tuberculosis Foundation; after long-term mechanical loading is greatest
p. 311–316. if loading is separated into short bouts. Journal
Roberts CA. 2002. The antiquity of leprosy in of Bone and Mineral Research 17:1545–1554.
Britain: the skeletal evidence. In: Roberts CA, Robling AG, Stout SD. 2003. Histomorphology,
Lewis ME, Manchester K, editors. The Past and geometry, and mechanical loading in past
Present of Leprosy: Archaeological, Historical, populations. In: Agarwal SC, Stout SD, editors.
Palaeopathological and Clinical Approaches. Bone Loss and Osteoporosis: An
British Archaeological Reports, International Anthropological Perspective. New York, NY:
Series, No. 1054; p. 213–222. Kluwer Academic Plenum; p. 189–205.
Roberts CA. 2007. A bioarchaeological study of Roche MB, Rowe GG. 1951. The incidence of
maxillary sinusitis. American Journal of separate neural arch and coincident bone
Physical Anthropology 133:792–807. variations. Journal of Bone and Joint Surgery
Roberts CA. 2009. Human Remains in 34-A:491–494.
Archaeology: A Handbook. York, UK: Council Rogers J, Dieppe P. 2003. Paleopathology of
for British Archaeology. osteoarthritis. In: Brandt KD, Doherty M,
Roberts CA. 2011. The bioarchaeology of leprosy Lohmander LS, editors. Osteoarthritis, Second
and tuberculosis: a comparative study of Edition. New York, NY: Oxford University
perceptions, stigma, diagnosis, and treatment. Press; p. 59–65.
In: Agarwal SC, Glencross BA, editors. Social Rogers J, Waldron T. 1995. A Field Guide to Joint
Bioarchaeology. Chichester, UK: Wiley- Disease in Archaeology. Chichester, UK: John
Blackwell; p. 254–283. Wiley & Sons.
Roberts CA, Boylston A, Buckley L, Chamberlain Rojas-Sepúlveda C, Ardagna Y, Dutour O. 2008.
AC, Murphy EM. 1998. Rib lesions and Paleoepidemiology of vertebral degenerative
tuberculosis: the palaeopathological evidence. disease in a Pre-Columbian Muisca series from
Tubercle and Lung Disease 79:55–60. Colombia. American Journal of Physical
Roberts CA, Buikstra JE. 2003. The Bioarchaeology Anthropology 135:416–430.
of Tuberculosis. Gainesville, FL: University Roksandic M, Djuric M, Rakocevic Z, Seguin K.
Press of Florida. 2006. Interpersonal violence at Lepenski Vir
Roberts, CA, Buikstra, JE. 2007. The evidence of Mesolithic/Neolithic complex of the Iron Gates
tuberculosis in the eastern Mediterranean: past Gorge (Serbia–Romania). American Journal of
and current research, and future prospects. In: Physical Anthropology 129:339–348.
Faerman M, Horwitz LK, Kahana T, Zilberman Romero J. 1970. Dental mutilation, trephination,
U, editors. Faces from the Past: Diachronic and cranial deformation. In: Stewart TD, editor.
Patterns in the Biology of Human Populations Handbook of Middle American Indians, Volume
from the Eastern Mediterranean. British 9: Physical Anthropology. Austin, TX:
Archaeological Reports, International Series, University of Texas Press; p. 50–67.
No. 1603; p. 213–227. Ronnerman A. 1977. The effect of early loss of
Roberts CA, Grauer A. 2001. Commentary: bones, primary molars on tooth eruption and space
bodies and representivity in the archaeological conditions: a longitudinal study. Acta
record. International Journal of Epidemiology Odontologica Scandinavica 35:229–239.
30:109–110. Rose F. 2008. Intra-community variation in diet
Roberts CA, Lewis ME, Manchester K, editors. during the adoption of a new staple crop in the
2002. The Past and Present of Leprosy: Eastern Woodlands. American Antiquity
Archaeological, Historical, Palaeopathological 73:413–439.
and Clinical Approaches. British Archaeological Rose JC. 1977. Defective enamel histology
Reports, International Series, No. 1054. of prehistoric teeth from Illinois. American
Robling AG, Hinant FM, Burr DB, Turner CH. Journal of Physical Anthropology
2002. Improved bone structure and strength 46:439–446.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
550 References

Rose JC. 1985. Gone to a Better Land: A Biohistory Rösing FW, Schwidetzky I. 1977. Vergleischend-
of a Rural Black Cemetery in the Post- statistische untersuchungen zur anthropologie
Reconstruction South. Arkansas des frohen Mittelalters (500–1000 n.d.Z). Homo
Archaeological Survey Research Series, No. 25. 28:65–115.
Rose JC, Anton SC, Aufderheide AC, et al. Rösing FW, Schwidetzky I. 1979. Comparative
1991. Skeletal Database Committee statistical studies on the physical anthropology
Recommendations. Detroit, Michigan: of the Early Medieval period (500–1000 A.D.).
Paleopathology Association. Journal of Human Evolution 8:755–758.
Rose JC, Armelagos GJ, Lallo JW. 1978. Rösing FW, Schwidetzky I. 1984. Comparative
Histological enamel indicator of childhood statistical studies on the physical anthropology
stress in prehistoric skeletal samples. American of the Late Medieval period (1000–1500).
Journal of Physical Anthropology 49:511–516. Journal of Human Evolution 13:325–329.
Rose JC, Armelagos GJ, Perry LS. 1993. Dental Ross AH, Ubelaker DH. 2010. A morphometric
anthropology of the Nile Valley. In: Davies approach to Taino biological distance in the
WV, Walker R, editor. Biological Anthropology Caribbean. In: Fitzpatrick SM, Ross AH, editors.
and the Study of Ancient Egypt. London, UK: Island Shores, Distant Pasts: Archaeological
British Museum Press; p. 61–74. and Biological Approaches to the Pre-
Rose JC, Burnett BA, Blauer MW, Nassaney MS. Columbian Settlement of the Caribbean.
1984. Paleopathology and the origins of maize Gainesville, FL: University Press of Florida;
agriculture in the lower Mississippi valley. In: p. 109–126.
Cohen MN, Armelagos GJ, editors. Ross-Stallings N. 1989. Treponemal syndrome at
Paleopathology at the Origins of Agriculture. the Austin site (22-Tu-549): a preliminary
Orlando, FL: Academic Press; p. 393–424. report. Mississippi Archaeology 24:1–16.
Rose JC, Marks MK, Tieszen LL. 1991. Rothhammer F, Silva C. 1990. Craniometrical
Bioarchaeology and subsistence in the central variation among South American prehistoric
and lower portions of the Mississippi Valley. populations: climatic, altitudinal,
In: Powell ML, Bridges PS, Mires AMW, editors. chronological, and geographic contributions.
What Mean These Bones? Studies in American Journal of Physical Anthropology
Southeastern Bioarchaeology. Tuscaloosa, AL: 82:9–17.
University of Alabama Press; p. 7–21. Rothschild BM, Rothschild C. 1995. Treponemal
Rose JC, Ungar PS. 1998. Gross dental wear and disease revisited: skeletal discriminators for
dental microwear in historical perspective. In: Alt yaws, bejel, and venereal syphilis. Clinical
KW, Rösing FW, Teschler-Nicola M, editors. Infectious Diseases 20:1402–1408.
Dental Anthropology: Fundamentals, Limits, and Rowe NH. 1982. Dental caries. In: Steele PF,
Prospects. Vienna, Austria: Springer; p. 349–386. editor. Dimensions of Dental Health, Third
Rosenberg AE. 1994. Skeletal system and soft Edition. Philadelphia, PA: Lea and Febiger;
tissue tumors. In: Cotran RS, Kumar V, Robbins p. 209–237.
SL, editors. Robbins Pathologic Basis of Roy TA, Ruff CB, Plato CC. 1994. Hand dominance
Disease, Fifth Edition. Philadelphia, PA: W.B. and bilateral asymmetry in the structure of the
Saunders; p. 1213–1271. second metacarpal. American Journal of
Rosenthal HL. 1981. Content of stable strontium in Physical Anthropology 94:203–211.
man and animal biota. In: Skoryna SC, editor. Rubini M, Zaio P. 2009. Lepromatous leprosy in an
Handbook of Stable Strontium. New York, NY: early Mediaeval cemetery in central Italy
Plenum Press; p. 1213–1271. (Morrione, Campochiaro, Molise, 6th–7th
Rösing FW. 1986. Kith or kin? On the feasibility of century AD). Journal of Archaeological Science
kinship reconstruction in skeletons. In: David 36:2771–2779.
AR, editor. Science in Egyptology. Manchester, Ruff CB. 1980. Age differences in craniofacial
UK: Manchester University Press; p. 223–237. dimensions among adults from Indian Knoll,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 551

Kentucky. American Journal of Physical Ruff CB. 2006. Environmental influences on


Anthropology 53:101–108. skeletal morphology. In: Ubelaker DH, editor.
Ruff CB. 1981. A reassessment of demographic Handbook of North American Indians:
estimates for Pecos Pueblo. American Journal Environment, Origins, and Population.
of Physical Anthropology 54:147–151. Washington, DC: Smithsonian Institution
Ruff CB. 1984. Allometry between length and Press; p. 685–693.
cross-sectional dimensions of the femur and Ruff CB. 2008. Biomechanical analyses of
tibia in Homo sapiens sapiens. American archaeological human skeletons. In: Katzenberg
Journal of Physical Anthropology 65:347–358. MA, Saunders SR, editors. Biological
Ruff CB. 1987. Sexual dimorphism in human Anthropology of the Human Skeleton, Second
lower limb bone structure: relationship to Edition. Hoboken, NJ: Wiley-Liss; p. 183–206.
subsistence strategy and sexual division of Ruff CB. 2010a. Bioarchaeology and Population
labor. Journal of Human Evolution History of Asia. American Journal of Physical
16:391–416. Anthropology. www.wiley.com/bw/wileyvi.
Ruff CB. 1989. New approaches to structural asp?ref=0002–9483&site=1.
evolution of limb bones in primates. Folia Ruff CB. 2010b. Structural analysis of postcranial
Primatologica 53:142–159. skeletal remains. In: Morgan ME, editor. Pecos
Ruff CB. 1991. Aging and Osteoporosis in Native Pueblo Revisited: the Biological and Social
Americans from Pecos Pueblo, New Mexico: Context. Cambridge, MA: Papers of the
Behavioral and Biomechanical Effects. New Peabody Museum of Archaeology and
York, NY: Garland Publishing. Ethnology, Harvard University, 85; p. 93–108.
Ruff C.B. 1992. Biomechanical analyses of Ruff CB, Garofalo E, Holmes MA. 2013.
archaeological human skeletal samples. In: Interpreting skeletal growth in the past from a
Saunders SR, Katzenberg MA, editors. The functional and physiological perspective.
Skeletal Biology of Past Peoples: Advances in American Journal of Physical Anthropology
Research Methods. New York, NY: Wiley-Liss; 150:29–37.
p. 37–58. Ruff CB, Hayes WC. 1982. Subperiosteal
Ruff CB. 1994a. Morphological adaptation to expansion and cortical remodeling of the
climate in modern and fossil hominids. human femur and tibia with aging. Science
Yearbook of Physical Anthropology 217:945–948.
37:65–107. Ruff CB, Hayes WC. 1983a. Cross-sectional
Ruff CB. 1994b. Biomechanical analysis of geometry of Pecos Pueblo femora and tibiae – a
northern and southern Plains femora: biomechanical investigation: I. Method and
behavioral implications. In: Owsley DW, Jantz general patterns of variation. American
RL, editors. Skeletal Biology in the Great Journal of Physical Anthropology
Plains: Migration, Warfare, Health, and 60:359–381.
Subsistence. Washington, DC: Smithsonian Ruff CB, Hayes WC. 1983b. Cross-sectional
Institution Press; p. 235–245. geometry of Pecos Pueblo femora and tibiae – a
Ruff CB. 1999. Skeletal structure and behavioral biomechanical investigation: II. Sex, age, and
patterns of prehistoric Great Basin populations. side differences. American Journal of Physical
In: Hemphill BE, Larsen CS, editors. Anthropology 60:383–400.
Understanding Prehistoric Lifeways in the Ruff CB, Holt BH, Trinkaus E. 2006. Who’s afraid
Great Basin Wetlands: Bioarchaeological of the big bad Wolff?: “Wolff’s law” and bone
Reconstruction and Interpretation. Salt Lake functional adaptation. American Journal of
City, UT: University of Utah Press; p. 290–320. Physical Anthropology 129:484–498.
Ruff CB. 2003. Growth in bone strength, body size, Ruff CB, Jones HH. 1981. Bilateral asymmetry in
and muscle size in a juvenile longitudinal cortical bone of the humerus and tibiae – sex
sample. Bone 33:317–329. and age factors. Human Biology 53:69–86.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
552 References

Ruff CB, Larsen CS. 1990. Postcranial Rugg-Gunn AJ, Al-Mohammadi SM, Butler TJ.
biomechanical adaptations to subsistence 2000. Malnutrition and developmental defects
strategy changes on the Georgia coast. In: of enamel in 2- to 6-year-old Saudi boys.
Larsen CS, editor. The Archaeology of Mission Caries Research 32:181–192.
Santa Catalina de Guale: 2. Biocultural Rugg-Gunn AJ, Hackett AF, Appleton DR, Jenkins
Interpretations of a Population in Transition. GN, Eastoe JE. 1984. Relationship between
Anthropological Papers of the American dietary habits and caries increment assessed
Museum of Natural History, No. 68; p. 94–120. over two years in 405 English adolescent
Ruff CB, Larsen CS. 2001. Reconstructing behavior school children. Archives of Oral Biology
in Spanish Florida: the biomechanical 29:983–992.
evidence. In: Larsen CS, editor. Bioarchaeology Runestad JA, Ruff CB, Nieh JC, Thorington RW,
of Spanish Florida: The Impact of Colonialism. Teaford MF. 1993. Radiographic estimation of
Gainesville, FL: University Press of Florida; long bone cross-sectional geometric properties.
p. 113–145. American Journal of Physical Anthropology
Ruff CB, Larsen CS. 2014. Long bone structural 90:207–213.
analyses and the reconstruction of past Runia LT. 1987a. Strontium and calcium
mobility: a historical review. In: Carlson KJ, distribution in plants: effect on palaeodietary
Marchi D, editors. Reconstructing Mobility: studies. Journal of Archaeological Science
Environmental, Behavioral, and Morphological 14:599–608.
Determinants. New York, NY: Springer; Runia LT. 1987b. Analysis of bone from the
p. 13–29. Bronze Age site Bovenkarspel-Het Valkje, the
Ruff CB, Larsen CS, Hayes WC. 1984. Structural Netherlands: a preliminary report.
changes in the femur with the transition to Archaeometry 29:221–232.
agriculture on the Georgia coast. American Russell AL, Consolazio CF, White CL. 1961. Dental
Journal of Physical Anthropology 64:125–136. caries and nutrition in Eskimo scouts of the
Ruff CB, Leo FP. 1986. Use of computed Alaska National Guard. Journal of Dental
tomography in skeletal structural research. Research 40:594–603.
Yearbook of Physical Anthropology Russell KF, Choi I, Larsen CS. 1990. The
29:181–195. paleodemography of Santa Catalina de Guale.
Ruff CB, McHenry HM, Thackeray JF. 1999. Cross- In: Larsen CS, editor. The Archaeology of
sectional morphology of the SK 82 and 97 Mission Santa Catalina de Guale: 2.
proximal femur. American Journal of Physical Biocultural Interpretations of a Population in
Anthropology 109:509–521. Transition. Anthropological Papers of the
Ruff CB, Runestad JA. 1992. Primate limb bone American Museum of Natural History, No. 68;
structural adaptations. Annual Review of p. 36–49.
Anthropology 21:407–433. Russell MD. 1985. The supraorbital torus: “a most
Ruff CB, Scott WW, Liu AY-C. 1991. Articular and remarkable peculiarity.” Current Anthropology
diaphyseal remodeling of the proximal femur 26:337–360.
with changes in body mass in adults. American Russell N, Twiss KC, Orton DC, Demirergi A. 2013.
Journal of Physical Anthropology 86:397–413. More on the Çatalhöyük mammal remains. In:
Ruff CB, Trinkaus E, Walker A, Larsen CS. 1993. Hodder I, editor. Humans and Landscapes of
Postcranial robusticity in Homo. I: Temporal Çatalhöyük. Los Angeles, CA: Cotsen Institute
trends and mechanical interpretation. of Archaeology Press; p. 213–258.
American Journal of Physical Anthropology Russo CR, Lauretani F, Seeman E, et al. 2006.
91:21–53. Structural adaptations to bone loss in aging
Ruff CB, Walker A, Trinkaus E. 1994. Postcranial men and women. Bone 38:112–118.
robusticity in Homo. III: Ontogeny. American Ryan AS. 1976. Long bone growth in a prehistoric
Journal of Physical Anthropology 93:35–54. population from San Cristobal, New Mexico.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 553

Michigan Discussions in Anthropology Salo WL, Aufderheide AC, Buikstra J, Holcomb


2:55–75. TA. 1994. Identification of Mycobacterium
Ryan AS. 1981. Anterior dental microwear and its tuberculosis DNA in a pre-Columbian Peruvian
relationship to diet and feeding behavior in mummy. Proceedings of the National Academy
three African primates (Pan troglodytes of Sciences 91:2091–2094.
troglodytes, Gorilla gorilla gorilla, and Papio Salvadei L, Ricci F, Manzi G. 2001. Porotic
hamadryas). Primates 22:533–550. hyperostosis as a marker of health and
Ryan AS, Johanson DC. 1989. Anterior dental nutritional conditions during childhood:
microwear in Australopithecus afarensis: studies at the transition between Imperial
comparisons with human and nonhuman Rome and the early Middle Ages. American
primates. Journal of Human Evolution Journal of Human Biology 13:709–717.
18:235–268. Sandberg LG, Steckel RH. 1987. Heights and
Ryan TM, Milner GR. 2006. Osteological economic history: the Swedish case. Annals of
applications of high-resolution tomography: a Human Biology 14:101–110.
prehistoric arrow injury. Journal of Sandford MK. 1992. A reconsideration of trace
Archaeological Science 33:871–879. element analysis in prehistoric bone. In:
Rylander KA. 1994. Corn preparation among the Saunders SR, Katzenberg MA, editors. Skeletal
Basketmaker Anasazi: a scanning electron Biology of Past Peoples: Research Methods.
microscope study of Zea mays remains from New York, NY: Wiley-Liss; p. 79–103.
coprolites. In: Sobolik KD, editor. Sandford MK, editor. 1993a. Investigations of
Paleonutrition: The Diet and Health of Ancient Human Tissue: Chemical Analyses in
Prehistoric Americans. Carbondale, IL: Anthropology. Langhorne, PA: Gordon and
Southern Illinois University at Carbondale, Breach Science Publishers.
Center for Archaeological Investigations, Sandford MK. 1993b. Understanding the biogenic-
Occasional Paper 22; p. 115–133. diagenetic continuum: interpreting elemental
Sadr K. 2003. The Neolithic of southern Africa. concentrations of archaeological bone. In:
Journal of African History 44:195–209. Sandford MK, editor. Investigations of Ancient
Sagne S. 1976. The jaws and teeth of a medieval Human Tissue: Chemical Analyses in
population in Southern Sweden: an Anthropology. Langhorne, PA: Gordon and
anthropological study of a skull material with Breach Scientific Publishers; p. 3–57.
special reference to attrition, size of jaws and Sandford MK, Bogdan G, Kissling GE. 2002.
teeth, and third-molar impaction. Ossa 3 Biological adaptation in the prehistoric
(Supplement 1). Caribbean: Osteology and bioarchaeology
Sahlins M. 1972. Stone Age Economics. Chicago, of the Tutu site. In: Righter E, editor. The
IL: Aldine Publishing. Tutu Archaeological Village Site: A
Sakashita R, Inoue M, Inoue N, Pan Q, Zhu H. Multidisciplinary Case Study in Human
1997. Dental disease in the Chinese Yin-Shang Adaptation. New York, NY: Routledge;
period with respect to relationships between p. 209–229.
citizens and slaves. American Journal of Sandford MK, Repke DB, Earle AL. 1988.
Physical Anthropology 103:401–408. Elemental analysis of human bone from
Sakellarakis JA, Sapouna-Sakellaraki E. 1991. Carthage: a pilot study. In: Humphrey JH,
Archanes. Athens, Greece: Ekdotike Athenon S.A. editor. The Circus and a Byzantine Cemetery at
Salib P. 1967. Trauma and disease of the post- Carthage, Volume 1. Ann Arbor, MI: University
cranial skeleton in ancient Egypt. In: Brothwell of Michigan Press; p. 285–296.
D, Sandison AT, editors. Diseases in Antiquity: Sandzen SC Jr. 1979. Atlas of Wrist and Hand
A Survey of the Diseases, Injuries, and Surgery Fractures. Littleton, MA: PSG Publishing.
of Early Populations. Springfield, IL: Charles Sankhyan AR, Weber GHJ. 2001. Evidence of
C. Thomas; p. 599–605 surgery in ancient India: trepanation at

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
554 References

Burzahom (Kashmir) over 4000 years ago. early childhood. Annals of Human Biology
International Journal of Osteoarchaeology 7:359–365.
11:375–380. Sauerwein E. 1974. Kariologie. Stuttgart,
Santley RS. 1980. Disembedded capitals Germany: G. Thieme.
reconsidered. American Antiquity 45: Saul FP. 1972. The Human Skeletal Remains of
132–145. Altar de Sacrificios. Harvard University, Papers
Santos AL, Roberts CA. 2001. A picture of of the Peabody Museum of Archaeology and
tuberculosis in young Portuguese people in the Ethnology, 63(2).
early 20th century: a multidisciplinary study of Saul FP, Saul JM. 1991. The Preclassic population
the skeletal and historical evidence. American of Cuello. In: Hammond N, editor. Cuello: An
Journal of Physical Anthropology 115: Early Maya Community in Belize. Cambridge,
38–49. UK: Cambridge University Press; p. 134–158.
Santos RV, Coimbra CEA Jr. 1999. Hardships of Saunders SR. 1978. The Development and
contact: enamel hypoplasia in Tupí-Mondé Distribution of Discontinuous Morphological
Amerindians from the Brazilian Amazonia. Variation of the Human Infracranial Skeleton.
American Journal of Physical Anthropology Archaeological Survey of Canada, Mercury
109:111–127. Series Paper, No. 81.
Santure SK. 1990. Social conflict. In: Santure SK, Saunders SR. 1989. Nonmetric skeletal variation.
Harn AD, Esarey D, editors. Archaeological In: İşcan MY, Kennedy KAR, editors.
Investigations at the Morton Village and Norris Reconstruction of Life from the Skeleton. New
Farms 36 Cemetery. Illinois State Museum York, NY: Alan R. Liss; p. 95–108.
Reports of Investigations, No. 45; p. 154–159. Saunders SR. 1991. Sex determination, stature and
Sardi ML, Beguelin M. 2011. Skeletal size and shape variation of the limb bones. In:
differentiation at the southernmost frontier of Pfeiffer S, Williamson RF, editors. Snake Hill:
Andean agriculture. In: Pinhasi R, Stock JT, An Investigation of a Military Cemetery from
editors. Human Bioarchaeology of the the War of 1812. Toronto, ON: Dundurn Press;
Transition to Agriculture. Chichester, UK: p. 176–187.
Wiley-Blackwell; p. 429–450. Saunders SR. 2008. Juvenile skeletons and
Sardi ML, Novellino PS, Pucciarelli HM. 2006. growth-related studies. In: Katzenberg MA,
Craniofacial morphology in the Argentine Saunders SR, editors. Biological Anthropology
center-west: consequences of the transition to of the Human Skeleton, Second Edition.
food production. American Journal of Physical Hoboken, NJ: Wiley-Liss; p. 117–147.
Anthropology 130:333–343. Saunders SR, Herring DA, Boyce G. 1995. Can
Sardi ML, Rozzi R, Pucciarelli HM. 2004. The skeletal samples accurately represent the living
Neolithic transition in Europe and North populations they come from? The St. Thomas’
Africa. The functional craniology contribution. cemetery site, Belleville, Ontario. In: Grauer
Anthropologischer Anzeiger 62:129–145. AL, editor. Bodies of Evidence: Reconstructing
Sarnat BG, Schour I. 1942. Enamel hypoplasia History through Skeletal Analysis. New York,
(chronologic enamel aplasia) in relation to NY: Wiley-Liss; p. 69–80.
systemic disease: a chronologic, morphologic, Saunders SR, Herring A, Sawchuk L, et al. 2002.
and etiologic classification. Journal of the The health of the middle class: the St. Thomas’
American Dental Association 29:67–77. Anglican Church cemetery project. In: Steckel
Sattenspiel L, Harpending H. 1983. Stable RH, Rose JC, editors. The Backbone of History:
populations and skeletal age. American Health and Nutrition in the Western
Antiquity 48:489–498. Hemisphere. New York, NY: Cambridge
Satyanarayana K, Naidu AN, Rao BSN. 1980 University Press; p. 130–161.
Adolescent growth spurt among rural Indian Saunders SR, Hoppa RD. 1993. Growth deficit in
boys in relation to their nutritional status in survivors and non-survivors: biological

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 555

mortality bias in subadult skeletal samples. temporal differences in diet at Piedras Negras,
Yearbook of Physical Anthropology Guatemala. Latin American Antiquity
36:127–151. 18:85–104.
Saunders SR, Keenleyside A. 1999. Enamel Schermer SJ, Fisher AK, Hodges DC. 1994.
hypoplasia in a Canadian historic sample. Endemic treponematosis in prehistoric
American Journal of Human Biology western Iowa. In: Owsley DW, Jantz RL,
11:513–524. editors. Skeletal Biology in the Great Plains:
Saunders SR, Lazenby R. 1991. The Links that Migration, Warfare, Health, and Subsistence.
Bind: The Harvie Family Nineteenth Century Washington. DC: Smithsonian Institution
Burying Ground. Occasional Papers in Press; p. 109–121.
Northeastern Archaeology, No. 5. Dundas, ON: Scheuer L, Black S. 2000. Developmental Juvenile
Capetown Press. Osteology. Amsterdam, the Netherlands:
Saunders SR, Popovich F. 1978. A family study of Elsevier Academic Press.
two skeletal variants: atlas bridging and Scheuer L, Black S. 2004. The Juvenile Skeleton.
clinoid bridging. American Journal of Physical Amsterdam, the Netherlands: Elsevier
Anthropology 49:193–204. Academic Press.
Saunders SR, Ramsden PG, Herring DA. 1992. Schillaci MA, Irish JD, Wood CCE. 2009. Further
Transformation and disease: precontact analysis of population history of ancient
Ontario Iroquoians. In: Verano JW, Ubelaker Egyptians. American Journal of Physical
DH, editors. Disease and Demography in the Anthropology 139:235–243.
Americas. Washington, DC: Smithsonian Schillaci MA, Nikitovic D, Akins NJ, Tripp L,
Institution Press; p. 117–125. Palkovich AM. 2011. Infant and juvenile
Savoye I, Loos R, Carels C, Derom C, Vlietinck R. growth in ancestral Pueblo Indians. American
1998. A genetic study of anteroposterior and Journal of Physical Anthropology
vertical facial proportions using model-fitting. 145:318–326.
Angle Orthodontist 68:467–470. Schillaci MA, Stojanowski CM. 2002.
Sawyer DR, Nwoku AL. 1985. Malnutrition and A reassessment of matrilocality in Chacoan
the oral health of children in Ogbomosho, culture. American Antiquity 67:343–356.
Nigeria. Journal of Dentistry for Children Schillaci MA, Stojanowski CM. 2003. Postmarital
52:141–145. residence and biological variation at Pueblo
Schaaffhausen H. 1858. Zur Kenntniss der ältesten Bonito. American Journal of Physical
Rassenschädel. Archiv für Anatomie, Anthropology 120:1–15.
Physiologie und wissenschaftliche Medicin Schillaci MA, Stojanowski CM. 2005. Craniometric
25:453–478. variation and population history of the
Schaafsma G, van Beresteyn ECH, Raymakers JA, prehistoric Tewa. American Journal of Physical
Duursma SA. 1987. Nutritional aspects of Anthropology 126:404–412.
osteoporosis. In: Bourne GH, editor. Nutrition Schimmer EL. 1979. De Romeinse waterputten te
and the Quality of Life. World Review of Velsen. Westerheem 28:109–116.
Nutrition and Dietetics 49; p. 121–159. Schindler DL. 1985. Anthropology in the Arctic:
Schaeffer AA. 1928. Spiral movement in man. a critique of racial typology and normative
Journal of Morphology and Physiology theory. Current Anthropology 26:475–500.
45:293–398. Schindler DL, Armelagos GJ, Bumsted MP. 1981.
Schepartz LA, Fox SC, Bourbou C, editors. 2009. Biocultural adaptation: new directions in
New Directions in the Skeletal Biology of Northeastern anthropology. In: Snow DR,
Greece. Athens, Greece: American School of editor. Foundations of Northeast Archaeology.
Classical Studies. New York, NY: Academic Press; p. 229–259.
Scherer AK, Wright LE, Yoder CJ. 2007. Schlecht SH. 2012. Understanding entheses:
Bioarchaeological evidence for social and bridging the gap between clinical and

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
556 References

anthropological perspectives. Anatomical and strontium levels in bone mineral. Journal


Record 295:1239–1251. of Human Evolution 14:515–525.
Schlecht SH, Pinto DC, Agnew AM, Stout SD. Schoeninger MJ. 1989. Reconstructing prehistoric
2012. The effects of disuse on the mechanical human diet. In: Price TD, editor. The Chemistry
properties of bone: what unloading tells us of Prehistoric Human Bone. Cambridge, UK:
about the adaptive nature of skeletal tissue. Cambridge University Press; p. 38–67.
American Journal of Physical Anthropology Schoeninger MJ. 1995a. Stable isotope studies in
149:599–605. human evolution. Evolutionary Anthropology
Schmidt BE, Schroder IW, editors. 2001. 4:83–98.
Anthropology of Violence and Conflict. London, Schoeninger MJ. 1995b. Dietary reconstruction in
UK: Routledge. the prehistoric Carson Desert: stable carbon
Schmidt CW. 2001. Dental microwear evidence for and nitrogen isotopic analysis. In: Larsen CS,
a dietary shift between two nonmaize-reliant Kelly RL, editors. Bioarchaeology of the
prehistoric human populations from Indiana. Stillwater Marsh: Prehistoric Human
American Journal of Physical Anthropology Adaptation in the Western Great Basin.
114:139–145. Anthropological Papers of the American
Schmitt D, Churchill SE, Hylander WL. 2003. Museum of Natural History, No. 77; p. 96–106.
Experimental evidence concerning spear use in Schoeninger MJ. 2009. Stable isotope evidence for
Neandertals and early modern humans. Journal the adoption of maize agriculture. Current
of Archaeological Science 30:103–114. Anthropology 50:633–640.
Schmorl G, Junghanns H. 1971. The Human Spine Schoeninger MJ. 2010. Diet reconstruction and
in Health and Disease, Second (American) ecology using stable isotope ratios. In: Larsen
Edition. New York, NY: Grune & Stratton. CS, editor. A Companion to Biological
Schmucker BJ. 1985. Dental attrition: a correlative Anthropology. Malden, MA: Wiley-Blackwell;
study of dietary and subsistence patterns in p. 445–464.
California and New Mexico Indians. In: Merbs Schoeninger MJ, DeNiro MJ. 1982. Carbon isotope
CF, Miller RJ, editors. Health and Disease in the ratios of apatite from fossil bone cannot be
Prehistoric Southwest. Arizona State used to reconstruct diets of animals. Nature
University Anthropological Research Papers, 297:577–578.
No. 34; p. 275–323. Schoeninger MJ, DeNiro MJ. 1984. Nitrogen and
Schneider KN, Blakeslee DJ. 1990. Evaluating carbon isotopic composition of bone collagen
residence patterns among prehistoric from marine and terrestrial animals.
populations: clues from dental enamel Geochimica et Cosmochimica Acta
composition. Human Biology 62:71–83. 48:625–639.
Schoeninger MJ. 1979. Diet and status at Schoeninger MJ, DeNiro MJ, Tauber H. 1983.
Chalcatzingo: some empirical and technical Stable nitrogen isotope ratios of bone collagen
aspects of strontium analysis. American reflect marine and terrestrial components of
Journal of Physical Anthropology 51:295–310. prehistoric human diet. Science
Schoeninger MJ. 1981. The agricultural 220:1381–1383.
“revolution”: its effect on human diet in Schoeninger MJ, Moore K. 1992. Bone stable
prehistoric Iran and Israel. Paleorient 7: isotope studies in archaeology. Journal of
73–91. World Prehistory 6:247–296.
Schoeninger MJ. 1982. Diet and the evolution of Schoeninger MJ, Peebles CS. 1981. Effect of
modern human form in the Middle East. mollusc eating on human bone strontium
American Journal of Physical Anthropology levels. Journal of Archaeological Science
58:37–52. 8:391–397.
Schoeninger MJ. 1985. Trophic level effects of Schoeninger MJ, Sattenspiel L, Schurr MR. 2000.
15 14
N/ N and 13C/12C ratios in bone collagen Transitions at Moundville: a question of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 557

collapse. In: Lambert PM, editor. Biological Schulting R, Richards MP. 2001. Dating women
Studies of Life in the Age of Agriculture: and becoming farmers: new palaeodietary and
A View from the Southeast. Tuscaloosa, AL: A.M.S. dating evidence from the Breton
University of Alabama Press; p. 63–77. Mesolithic cemeteries of Teviec and Hoedic.
Schoeninger MJ, van der Merwe NJ, Moore K, Lee- Journal of Anthropological Archaeology
Thorp J, Larsen CS. 1990. Decrease in diet 20:314–344.
quality between the prehistoric and contact Schultz M. 2001. Paleohistology of bone: a new
periods. In: Larsen CS, editor. The Archaeology approach to the study of ancient diseases.
of Mission Santa Catalina de Guale: 2. Yearbook of Physical Anthropology
Biocultural Interpretations of a Population in 44:106–147.
Transition. Anthropological Papers of the Schultz M. 2003. Light microscopic analysis in
American Museum of Natural History, No. 68; skeletal paleopathology. In: Ortner DJ, editor.
p. 78–93. Identification of Pathological Conditions in
Schour I. 1936. The neonatal line in the enamel Human Skeletal Remains. San Diego, CA:
and dentin of the human deciduous teeth and Academic Press; p. 73–107.
first permanent molar. Journal of the American Schultz M, Larsen CS, Kreutz K. 2001. Disease in
Dentistry Association 23:1946–1955. Spanish Florida: microscopy of porotic
Schreinder KE. 1935. Zur Osteologie der Lappen. hyperostosis and inferences about health. In:
Oslo, Norway: H. Aschehoug & Company. Larsen CS, editor. Bioarchaeology of Spanish
Schroeder HA, Tipton IH, Nason AP. 1972. Trace Florida: The Impact of Colonialism.
metals in man: strontium and barium. Journal Gainesville, FL: University Press of Florida;
of Chronic Diseases 25:491–517. p. 207–225.
Schuenemann VJ, Bos K, DeWitte S, et al. 2011. Schulz PD. 1977. Task activity and anterior tooth
Targeted enrichment of ancient pathogens grooving in prehistoric California Indians.
yielding the pPCP1 plasmid of Yersinia pestis American Journal of Physical Anthropology
from victims of the Black Death. Proceedings of 46:87–91.
the US National Museum 108:746–752. Schumacher GH, Ivankievicz D, Fanghnel J. 1979.
Schuenemann VJ, Sign P, Mendum TA, et al. The role of function in the formation of the
2013. Genome-wide comparison of medieval skull. Acta Morphologica Academiae
and modern Mycobacterium leprae. Science Scientiarum Hungaricae 27:53–56.
341:179–183. Schurr MR. 1992. Isotopic and mortuary
Schulman AR. 1982. The battle scenes of the variability in a Middle Mississippian
Middle Kingdom. Journal of the Society for the population. American Antiquity 57:300–320.
Study of Egyptian Antiquities 12:165–183. Schurr MR. 1998. Using stable nitrogen-isotopes
Schulting R. 2006. Skeletal evidence and contexts to study weaning behavior in past populations.
of violence in the European Mesolithic and World Archaeology 30:327–342.
Neolithic. In: Gowland R, Knüsel C, editors. Schurr MR, Schoeninger MJ. 1995. Associations
Social Archaeology of Funerary Remains. between agricultural intensification and social
Oxford, UK: Oxbow Books; p. 224–237. complexity: an example from the prehistoric
Schulting R. 2011. Mesolithic–Neolithic Ohio Valley. Journal of Anthropological
transitions: an isotopic tour through Europe. Archaeology 14:315–339.
In: Pinhasi R, Stock JT, editors. Human Schurr MR, Powell ML. 2005. The role of changing
Bioarchaeology of the Transition to Agriculture. childhood diets in the prehistoric evolution of
Chichester, UK: Wiley-Blackwell; p. 17–41. food production: an isotopic assessment.
Schulting R, Fibiger L, editors. 2012. Sticks, American Journal of Physical Anthropology
Stones, and Broken Bones: Neolithic Violence 126:278–294.
in a European Perspective. Oxford, UK: Oxford Schurr TG. 2004. The peopling of the New World.
University Press. Annual Review of Anthropology 33:551–583.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
558 References

Schurr TG, Ballinger SW, Gan YY, et al. 1990. Schwidetzky I. 1967. Vergleichend-Statistische
Amerindian mitochondrial DNAs have rare Untersuchungen zur Anthropologie des
Asian mutations at high frequencies, Neolithikums: Ergebnisse der Penrose-
suggesting they derived from four primary Analyse. Das Gesamtmaterial. Homo
maternal lineages. American Journal of Human 18:174–198.
Genetics 46:613–623. Schwidetzky I. 1972. Vergleichend-Statistische
Schutkowski H, Herrmann B, Wiedemann F, Untersuchungen zur Anthropologie der
Bocherens H, Grupe G. 1999. Diet, status and Eisenzeit. Homo 23:245–272.
decomposition of Weingarten: trace element Schwidetzky I, Rösing FW. 1975. Vergleischend-
and isotope analyses on early Mediaeval Statistische Untersuchungen zur
skeletal material. Journal of Archaeological Anthropologie der Romerzeit (0–500 v.Z.).
Science 26:675–685. Homo 26:193–218.
Schwarcz HP, Dupras TL, Fairgrieve SI. 1999. Schwidetzky I, Rösing FW. 1982. European
15
N enrichment in the Sahara: in search of a populations of the high and late Medieval
global relationship. Journal of Archaeological period (1000–1500): comparative statistical
Science 26:629–636. studies on historical anthropology.
Schwarcz HP, Melbye J, Katzenberg MA, Knyf M. Humanbiologie (Budapest) 10:39–47.
1985. Stable isotopes in human skeletons of Schwitalla AW, Jones TL, Pilloud MA, Codding BF,
southern Ontario: reconstructing paleodiet. Wiberg RS. 2014. Violence among foragers: the
Journal of Archaeological Science 12:187–206. bioarchaeological record from central
Schwarcz HP, Schoeninger MJ. 1991. Stable California. Journal of Anthropological
isotope analyses in human nutritional ecology. Archaeology 33:66–83.
Yearbook of Physical Anthropology Sciulli PW. 1979. Size and morphology of the
34:283–321. permanent dentition in prehistoric Ohio Valley
Schwarcz HP, Schoeninger MJ. 2011. Stable Amerindians. American Journal of Physical
isotopes of carbon and nitrogen as tracers for Anthropology 50:615–628.
paleo-diet reconstruction. In: Baskaran M, Sciulli PW. 1990. Cranial metric and discrete trait
editor. Handbook of Environmental Isotope variation and biological differentiation in the
Geochemistry: Advances in Isotope terminal Late Archaic of Ohio: the Duff site
Geochemistry. Berlin, Germany: Springer- cemetery. American Journal of Physical
Verlag; p. 725–742. Anthropology 82:19–29.
Schwartz JH. 2006. Skeleton Keys: An Sciulli PW. 1997. Dental evolution in prehistoric
Introduction to Human Skeletal Morphology, Native Americans of the Ohio Valley area.
Development, and Analysis. New York, NY: I. Wear and pathology. International Journal of
Oxford University Press. Osteoarchaeology 7:507–524.
Schwartz JH, Brauer J, Gordon-Larsen P. 1995. Sciulli PW. 1998. Evolution of the dentition in
Tigaran (Point Hope, Alaska) tooth drilling. prehistoric Ohio Valley Native Americans. II.
American Journal of Physical Anthropology Morphology of the deciduous dentition.
97:77–82. American Journal of Physical Anthropology
Schwartz JH, Houghton F, Macchiarelli R, 106:189–206.
Bondioli L. 2010. Skeletal remains from Punic Sciulli PW, Blatt SH. 2008. Evaluation of juvenile
Carthage do not support systematic sacrifice of stature and body mass prediction. American
infants. PLoS ONE 5:e9177, 1–12. Journal of Physical Anthropology
Schwartz JH, Houghton FD, Bondioli L, 136:387–393.
Macchiarelli R. 2012. Bones, teeth, and Sciulli PW, Carlisle R. 1977. Analysis of the
estimating age of perinates: Carthaginian dentition from three western Pennsylvania
infant sacrifice revisited. Antiquity Late Woodland sites. II. Wear and pathology.
86:738–745. Pennsylvania Archaeologist 47:53–59.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 559

Sciulli PW, Doyle WJ, Kelley C, Siegel P, Siegel MI. Scott GR, Halffman CM, Pedersen PO. 1992.
1979. The interaction of stressors in the Dental conditions of Medieval Norsemen in the
induction of increased levels of fluctuating North Atlantic. Acta Archaeologica
asymmetry in the laboratory rat. American 62:183–207.
Journal of Physical Anthropology 50:279–284. Scott GR, Poulson SR. 2012. Stable carbon and
Sciulli PW, Oberly J. 2002. Native Americans in nitrogen isotopes of human dental calculus: a
eastern North America: the southern Great potentially new non-destructive proxy for
Lakes and upper Ohio Valley. In: Steckel RH, paleodietary analysis. Journal of
Rose JC, editors. The Backbone of History: Archaeological Science 39:1388–1393.
Health and Nutrition in the Western Scott GR, Turner CG II. 1988. Dental anthropology.
Hemisphere. New York, NY: Cambridge Annual Review of Anthropology 17:99–126.
University Press; p. 440–480. Scott GR, Turner CG II. 1997. The Anthropology of
Sciulli PW, Schneider KN, Mahaney MC. 1984. Modern Human Teeth: Dental Morphology and
Morphological variation of the permanent Its Variation in Recent Human Populations.
dentition in prehistoric Ohio. Anthropologie Cambridge, UK: Cambridge University Press.
22:211–215. Scott GR, Turner CG II. 2006. Dentition. In:
Scollard DM, Adams LB, Gillis TP, et al. 2006. The Ubelaker D, editor. Handbook of North
continuing challenges of leprosy. Clinical American Indians: 3. Environment, Origins,
Microbiology Reviews 19:338–381. and Population. Washington, DC: Smithsonian
Scott A, Whittaker J, Green M, Hillson S. 1991. Institution Press; p. 645–660.
Graphical modeling of archaeological data. In: Scott GR, Winn JR. 2011. Dental chipping:
Rahtz S, Lockyear K, editors. Computer contrasting patterns of microtrauma in Inuit
Applications and Quantitative Methods in and European populations. International
Archaeology 1990. British Archaeological Journal of Osteoarchaeology 21:723–731.
Reports, International Series, No. 565; Scott RM, Buckley HR. 2010. Biocultural
p. 111–116. interpretations of trauma in two prehistoric
Scott D, Willey P. 1997. Little Bighorn: human Pacific Island populations from Papua New
remains from the Custer National Cemetery. In: Guinea and the Solomon Islands. American
Poirier DA, Bellantoni NF, editors. In Journal of Physical Anthropology
Remembrance: Archaeology and Death. 142:509–518.
Westport, CT: Bergin & Garvey; p. 155–171. Scott RS, Teaford MF, Ungar PS. 2012. Dental
Scott DD, Willey P, Conner MA. 2002. They Died microwear texture and anthropoid diet.
with Custer: Soldiers’ Bones from the Battle of American Journal of Physical Anthropology
the Little Bighorn. Norman, OK: University of 147:551–579.
Oklahoma Press. Scott S, Duncan CJ. 2001. The Biology of Plagues.
Scott EC. 1979. Increase in tooth size in prehistoric Cambridge, UK: Cambridge University Press.
coastal Peru, 10,000 B.P.–1,000 B.P. American Scrimshaw NS. 2003. Historical concepts of
Journal of Physical Anthropology 50:251–258. interactions, synergism and antagonism
Scott GR. 1994. Teeth and prehistory on Kodiak between nutrition and infection. Journal of
Island. In: Bray TL, Killion TW, editors. Nutrition 133:316S–321S.
Reckoning with the Dead: the Larsen Bay Scrimshaw NS. 2010. INCAP studies of nutrition
Repatriation and the Smithsonian Institution. and infection. Food and Nutrition Bulletin
Washington, DC: Smithsonian Institution 31(1):S4–S7.
Press; p. 67–74. Scrimshaw NS, Taylor CE, Gordon JE. 1968.
Scott GR. 2008. Dental morphology. In: Interaction of Nutrition and Infection. World
Katzenberg MA, Saunders SR, editors. Health Organization Monograph, No. 57.
Biological Anthropology of the Human Sealy J. 1986. Stable Carbon Isotopes and
Skeleton. Hoboken, NJ: Wiley-Liss; p. 265–298. Prehistoric Diets in the South-Western Cape

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
560 References

Province, South Africa. Cambridge pre-term, very-low-birthweight children.


Monographs in African Archaeology, No. 15; Journal of Dental Research 79:63–69.
British Archaeological Reports, International Seow WK. 1991. Enamel hypoplasia in the primary
Series, No. 293. dentition – a review. Journal of Dentistry for
Sealy JC, van der Merwe NJ. 1985. Isotope Children 58:441–452.
assessment of Holocene human diets in the Serrano Sanchez C. 1993. Funerary practices and
southwestern Cape, South Africa. Nature human sacrifice in Teotihuacan burials. In:
315:138–140. Berrin K, Pasztory E, editors. Teotihuacan: Art
Sealy JC, van der Merwe NJ. 1986. Isotope from the City of the Gods. New York, NY:
assessment and the seasonal-mobility Thames & Hudson; p. 108–115.
hypothesis in the southwestern Cape of South Service ER. 1962. Primitive Social Organization.
Africa. Current Anthropology 27:135–150. New York, NY: Random House.
Sealy JC, van der Merwe NJ, Lee Thorp JA, Service ER. 1966. The Hunters. Englewood Cliffs,
Lanham JL. 1987. Nitrogen isotopic ecology in NJ: Prentice-Hall.
southern Africa: implications for Sevastikoglou JA, Eriksson U, Larsson SE. 1969.
environmental and dietary tracing. Geochimica Skeletal changes of the amputation stump and
et Cosmochimica Acta 51:2707–2717. the femur on the amputated side. Acta
Sealy JC, van der Merwe NJ, Sillen A, Kruger FJ, Orthopaedica Scandinavica 40:624–633.
Krueger HW. 1991. 87Sr/86Sr as a dietary Shackelford LL. 2007. Regional variation in the
indicator in modern and archaeological bone. postcranial robusticity of late Upper Paleolithic
Journal of Archaeological Science 18:399–416. humans. American Journal of Physical
Seaman J, Leivesley S, Hogg C. 1984. Anthropology 133:655–668.
Epidemiology of Natural Disasters. Basel, Shahriaree H, Sajadi K, Rooholamini SA. 1979.
Switzerland: S. Karger. A family with spondylolisthesis. Journal of
Seckler D. 1980. “Malnutrition”: an intellectual Bone and Joint Surgery 61-A:1256–1258.
odyssey. Western Journal of Agricultural Shapiro HL. 1939. Migration and Environment.
Economics 5:219–227. New York, NY: Oxford University Press.
Seckler D. 1982. Small but healthy: a basic Sharma K, Corruccini RS, Potter RHY. 1986. Genetic
hypothesis in the theory, measurement and and environmental influences on bilateral dental
policy of malnutrition. In: Sukhatme PV, asymmetry in northwest Indian twins.
editor. New Concepts in Nutrition and Their International Journal of Anthropology 4:349–360.
Implications for Policy. Pune, India: Sharma L. 2001. Epidemiology of osteoarthritis.
Maharashtra Association for the Cultivation of In: Moskowitz RW, Howell DS, Altman RD,
Science Research Institute; p. 127–137. Buckwalter JA, Goldberg VM, editors.
Seidel JC. 1995. Activity-Induced Dental and Osteoarthritis: Diagnosis and Medical/Surgical
Osseous Conditions and the Division of Activity Management, Third Edition. Philadelphia, PA:
by Gender of the Protohistoric Seneca. PhD Saunders and Company; p. 3–27.
Dissertation, University of Connecticut, Shaw CN, Stock JT. 2009a. Habitual throwing and
Storrs, CT. swimming correspond with upper limb
Sejrsen B, Kjær I, Jakobsen J. 1996. Human palatal diaphyseal strength and shape in modern
growth evaluated on Medieval crania using human athletes. American Journal of Physical
nerve canal openings as references. American Anthropology 140:160–172.
Journal of Physical Anthropology 99:603–611. Shaw CN, Stock JT. 2009b. Intensity,
Seltzer CC. 1944. Racial Prehistory in the repetitiveness, and directionality of habitual
Southwest and the Hawikuh Zunis. New adolescent mobility patterns influence the
Haven, CT: Peabody Museum. tibial diaphysis morphology of athletes.
Seow W, Wan A. 2000. A controlled study of the American Journal of Physical Anthropology
morphometric changes in the primary dentin of 140:149–159.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 561

Shaw CN, Stock JT. 2013. Extreme mobility in the Silcox MT, Teaford MF. 2002. The diet of worms:
Late Pleistocene? Comparing limb an analysis of mole dental microwear. Journal
biomechanics among fossil Homo, varsity of Mammalogy 83:804–814.
athletes, and Holocene foragers. Journal of Sillen A. 1981. Strontium and diet at Hayonim
Human Evolution 64:242–249. Cave. American Journal of Physical
Shaw SR, Zernicke RF, Vailas AC, et al. 1987. Anthropology 56:131–137.
Mechanical, morphological and biochemical Sillen A. 1989. Diagenesis of the inorganic phase
adaptations of bone and muscle to hindlimb of cortical bone. In: Price TD, editor. The
suspension and exercise. Journal of Chemistry of Prehistoric Bone. Cambridge, UK:
Biomechanics 20:225–234. Cambridge University Press; p. 211–229.
Sheiham A. 2001. Dietary effects on dental Sillen A. 1992. Strontium – calcium ratios (Sr/Ca)
diseases. Public Health Nutrition 4(2b): of Australopithecus robustus and associated
569–591. fauna from Swartkrans. Journal of Human
Sheridan SG, Mittler DM, Van Gerven DP, Covert Evolution 23:495–516.
HH. 1991. Biomechanical association of dental Sillen A, Kavanagh M. 1982. Strontium and
and temporomandibular pathology in a paleodietary research: a review. Yearbook of
Medieval Nubian population. American Physical Anthropology 25:67–90.
Journal of Physical Anthropology 85:201–205. Sillen A, Sealy JC, van der Merwe NJ. 1989.
Shimada I, Shinoda K, Farnum J, Corruccini R, Chemistry and paleodietary research: no
Watanabe H. 2004. An integrated analysis of more easy answers. American Antiquity
pre-hispanic mortuary practices: a Middle 54:504–512.
Sicán case study. Current Anthropology Sillen A, Smith P. 1984. Weaning patterns are
45:369–402. reflected in strontium–calcium ratios of
Shimada I, Shinoda KI, Bourget S, Alva W, Uceda juvenile skeletons. Journal of Archaeological
S. 2005. mtDNA analysis of Mochica and Sicán Science 11:237–245.
populations of pre-Hispanic Peru. In: Reed DM, Silman AJ. 2003. Risk factors for Colles’ fracture
editor. Biomolecular Archaeology: Genetic in men and women: results from the European
Approaches to the Past. Southern Illinois Prospective Osteoporosis Study. Osteoporosis
University Center for Archaeological International 14:213–218.
Investigations Occasional Paper, No. 32; Simkin A, Leichter I, Swissa A, Samueloff S. 1989.
p. 61–92. The effect of swimming activity on bone
Shrewsbury JFD. 1970. A History of Bubonic architecture in growing rats. Journal of
Plague in the British Isles. Cambridge, UK: Biomechanics 22:845–851.
Cambridge University Press. Simmons DJ. 1985. Options for bone aging with
Sibley LM, Armelagos GJ, Van Gerven DP. 1992. the microscope. Yearbook of Physical
Obstetric dimensions of the true pelvis in a Anthropology 28:249–263.
Medieval population from Sudanese Nubia. Simms SR. 2010. Farmers, foragers, and adaptive
American Journal of Physical Anthropology diversity. In: Hemphill BE, Larsen CS, editors.
89:421–430. Prehistoric Lifeways in the Great Basin
Siegel MI, Doyle W, Kelly C. 1977. Heat stress, Wetlands: Bioarchaeological Reconstruction
fluctuating asymmetry and prenatal selection and Interpretation, Paperback edition. Salt
in the laboratory rat. American Journal of Lake City, UT: University of Utah Press;
Physical Anthropology 46:121–126. p. 21–54.
Siegel MI, Mooney MP. 1987. Perinatal stress and Simon JJK. 1992. Mortuary practices of the Late
increased fluctuating asymmetry of dental Kachemak Tradition in south central Alaska:
calcium in the laboratory rat. American a perspective from the Crag Point site,
Journal of Physical Anthropology Kodiak Island. Arctic Anthropology
73:267–270. 29:130–149.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
562 References

Simon JJK, Steffian AF. 1994. Cannibalism or Sjøvold T. 1984. A report on the heritability of
complex mortuary behavior? An analysis of some cranial measurements and non-metric
patterned variability in the treatment of human traits. In: van Vark GN, Howells WW, editors.
remains from the Kachemak tradition of Multivariate Statistical Methods in Physical
Kodiak Island. In: Bray TL, Killion TW, editors. Anthropology. Dordrecht, the Netherlands:
Reckoning with the Dead: The Larsen Bay D. Reidel Publishing Company; p. 223–246.
Repatriation and the Smithsonian Institution. Sjøvold T. 1995. Testing assumptions for skeletal
Washington, DC: Smithsonian Institution studies by means of identified skulls from
Press; p. 75–100. Hallstatt, Austria. In: Saunders SR, Herring A,
Simpson AHRW. 1985. The blood supply of the editors. Grave Reflections: Portraying the Past
periosteum. Journal of Anatomy 140:697–704. through Cemetery Studies. Toronto, ON:
Simpson JE. 1983 [1876]. Report of the Canadian Scholars’ Press; p. 241–281.
Explorations across the Great Basin of the Sládek V, Berner M, Sailer R. 2006a. Mobility in
Territory of Utah. Reno, NV: University of central European Late Eneolithic and Early
Nevada Press. Bronze Age. American Journal of Physical
Simpson SW. 1999. Reconstructing patterns of Anthropology 130:320–332.
growth disruption from enamel microstructure. Sládek V, Berner M, Sailer R. 2006b. Mobility in
In: Hoppa RD, FitzGerald CM, editors. Human Central European Late Eneolithic and Early
Growth in the Past: Studies from Bones and Bronze Age: tibial cross-sectional geometry.
Teeth. Cambridge, UK: Cambridge University Journal of Archaeological Science 33:470–482.
Press; p. 241–263. Sládek V, Berner M, Sosna D, Sailer R. 2007.
Simpson SW. 2001. Patterns of growth disruption Human manipulative behavior in the central
in La Florida: evidence from enamel European Late Eneolithic and Early Bronze
microstructure. In: Larsen CS, editor. Age: humeral bilateral asymmetry. American
Bioarchaeology of Spanish Florida: The Impact Journal of Physical Anthropology
of Colonialism. Gainesville, FL: University 133:669–681.
Press of Florida; p. 146–180. Šlaus M. 2000. Biocultural analysis of sex
Simpson SW, Hutchinson DL, Larsen CS. 1990. differences in mortality profiles and stress
Coping with stress: tooth size, dental defects, levels in the Late Medieval population from
and age-at-death. In: Larsen CS, editor. The Nova Raca, Croatia. American Journal of
Archaeology of Mission Santa Catalina de Physical Anthropology 111:193–209.
Guale: 2. Biocultural Interpretations of a Šlaus M, Novak M, Bedić Ž, Strinović D. 2012.
Population in Transition. Anthropological Bone fractures as indicators of intentional
Papers of the American Museum of Natural violence in the Eastern Adriatic from the
History, No. 68; p. 66–77. Antique to the Late Medieval period (2nd–16th
Sirirungrojying S, Kerdpon D. 1999. Relationship century AD). American Journal of Physical
between oral tori and temporomandibular Anthropology 149:26–38.
disorders. International Dental Journal Šlaus M, Novak M, Vyroubal V, Bedić Z. 2010. The
49:101–104. harsh life on the 15th century Croatia–Ottoman
Sjøvold T. 1973. Occurrence of minor non- empire military border: analyzing and
metrical variants in the skeleton and their identifying the reasons for the massacre at
quantitative treatment for population Cepin. American Journal of Physical
comparisons. Homo 24:204–233. Anthropology 141:358–372.
Sjøvold T. 1977. Non-metrical divergence Sledzik PS, Moore-Janssen PH. 1991. Dental
between skeletal populations: the theoretical disease in nineteenth century military skeletal
foundation and biological importance of C.A.B. samples. In: Kelley MA, Larsen CS, editors.
Smith’s Mean Measure of Divergence. Ossa Advances in Dental Anthropology. New York,
4(Supplement 1). NY: Wiley-Liss; p. 215–224.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 563

Sledzik PS, Sandberg LG. 2002. The effects of Smith GE, Jones FW. 1910. The Archaeological
nineteenth-century military service on health. Survey of Nubia, Report for 1907–1908,
In: Steckel RH, Rose JC, editors. The Backbone Volume II: Report on the Human Remains.
of History: Health and Nutrition in the Western Cairo, Egypt: National Printing Company.
Hemisphere. New York, NY: Cambridge Smith MJ. 2014. The war to begin all wars?
University Press; p. 185–207. Contextualizing violence in Neolithic Britain.
Sloan B, Kulwin DR, Kersten RC. 1999. Scurvy In: Knüsel C, Smith MJ, editors. The Routledge
causing bilateral orbital hemorrhage. Archives Handbook of the Bioarchaeology of Human
of Ophthalmology 117:842–843. Conflict. London, UK: Routledge;
Slots J. 2004. Update on human cytomegalovirus p. 109–126.
in destructive periodontal disease. Oral Smith MO. 1982. Patterns of Association between
Microbiology and Immunology 19:217–223. Oral Health Status and Subsistence: A Study of
Slovak NM, Paytan A, Wiegand BA. 2009. Aboriginal Skeletal Populations from the
Reconstructing Middle Horizon mobility Tennessee Valley Area. PhD Dissertation,
patterns on the coast of Peru through strontium University of Tennessee, Knoxville, TN.
isotope analysis. Journal of Archaeological Smith MO. 1993. A probable case of decapitation
Science 36:157–165. at the Late Archaic Robinson site (40SM4),
Smith BD. 1986. The archaeology of the Smith County, Tennessee. Tennessee
Southeastern United States: from Dalton to De Anthropologist 18:131–142.
Soto, 10,500–500 B.P. In: Wendorf F, Close AE, Smith MO. 1995. Scalping in the Archaic period:
editors. Advances in World Archaeology. evidence from the western Tennessee Valley.
Orlando, FL: Academic Press; p. 1–92. Southeastern Archaeology 14:60–68.
Smith BD. 1989. Origins of agriculture in eastern Smith MO. 1996. Biocultural inquiry into Archaic
North America. Science 246:1566–1571. period populations of the Southeast: trauma
Smith BD. 1998. The Emergence of Agriculture. and occupational stress. In: Sassaman KE,
New York, NY: Scientific American Library. Anderson DG, editors. Archaeology of the Mid-
Smith BD, Yarnell RA. 2009. Indigenous formation Holocene Southeast. Gainesville, FL: University
of an indigenous crop complex in eastern Press of Florida; p. 134–154.
North America at 3800 BP. Proceedings of the Smith MO. 1997. Osteological indications of
National Academy of Sciences 106:6561–6566. warfare in the Archaic period of the western
Smith BH. 1984. Patterns of molar wear in hunter- Tennessee Valley. In: Martin D, Frayer D,
gatherers and agriculturalists. American editors. Troubled Times: Violence and Warfare
Journal of Physical Anthropology 63:39–56. in the Past. Langhorne, PA: Gordon and
Smith BH. 1991. Standards of human tooth Breach; p. 241–265.
formation and dental age assessment. In: Smith MO. 2003. Beyond palisades: the nature and
Kelley MS, Larsen CS, editors. Advances in frequency of late prehistoric deliberate violent
Dental Anthropology. New York, NY: Wiley- trauma in the Chickamauga Reservoir of east
Liss; p. 143–168. Tennessee. American Journal of Physical
Smith BH, Garn SM, Cole PE. 1982. Problems of Anthropology 121:303–318.
sampling and inference in the study of Smith MO. 2008. Adding insult to injury:
fluctuating dental asymmetry. American opportunistic treponemal disease in a scalping
Journal of Physical Anthropology 58:281–289. survivor. International Journal of
Smith GE. 1908. The most ancient splints. British Osteoarchaeology 18:589–599.
Medical Journal 1:732–734. Smith MO, Betsinger TK, Williams LL. 2011.
Smith GE. 1910. The racial problem. In: Smith GE, Differential visibility of treponemal disease in
Jones FW, editors. The Archaeological Survey pre-Columbian stratified societies: does rank
of Nubia, Report for 1907–1908. Cairo, Egypt: matter? American Journal of Physical
National Printing Company; p. 15–36. Anthropology 144:185–195.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
564 References

Smith P. 1972. Diet and attrition in the Natufians. Kentucky Reports in Anthropology 4, No. 3,
American Journal of Physical Anthropology part II.
37:233–238. Snow CE. 1974. Early Hawaiians: An Initial Study
Smith P, Bar-Yosef O, Sillen A. 1984. of Skeletal Remains from Mokapu, Oahu.
Archaeological and skeletal evidence for Lexington, KY: University of Kentucky Press.
dietary change during the late Pleistocene/ Sobolik KD. 1994. Paleonutrition of the lower
Early Holocene in the Levant. In: Cohen MN, Pecos region of the Chihuahuan Desert. In:
Armelagos GJ, editors. Paleopathology at the Sobolik KD, editor. Paleonutrition: the Diet and
Origins of Agriculture. Orlando, FL: Academic Health of Prehistoric Americans. Southern
Press; p. 101–136. Illinois University at Carbondale, Center for
Smith P, Bloom RA, Berkowitz J. 1984. Diachronic Archaeological Investigations, Occasional
trends in humeral cortical thickness of Near Paper, No. 22; p. 247–264.
Eastern populations. Journal of Human Sofaer JR. 2006. The Body as Material Culture.
Evolution 13:603–611. Cambridge, UK: Cambridge University Press.
Smith P, Horwitz LK. 2007. Ancestors and Sofaer Derevenski JR. 2000. Sex differences in
inheritors: a bio-cultural perspective on the activity-related osseous change in the spine
transition to agro-pastoralism in the southern and the gendered division of labor at Ensay
Levant. In: Cohen MN, Crane-Kramer GMM, and Wharram Percy, UK. American Journal of
editors. Ancient Health: Skeletal Indicators of Physical Anthropology 111:333–354.
Agricultural and Economic Intensification. Sokal RR, Uytterschaut H. 1987. Cranial variation
Gainesville, FL: University Press of Florida; in European populations: a spatial
p. 207–222. autocorrelation study at three time periods.
Smith PG, Moss AR. 1994. Epidemiology of American Journal of Physical Anthropology
tuberculosis. In: Bloom BR, editor. 74:21–38.
Tuberculosis: Pathogenesis, Protection, and Sokal RR, Uytterschaut H, Rösing FW,
Control. Washington, DC: American Society Schwidetzky I. 1987. A classification of
for Microbiology Press; p. 47–59. European skulls from three time periods.
Smith RJ, Bailit HL. 1977. Problems and methods American Journal of Physical Anthropology
in research on the genetics of dental occlusion. 74:1–20.
Angle Orthodontist 47:65–77. Soler T, Calderón C. 2000. The prevalence of
Smith RW, Walker RR. 1964. Femoral expansion spondylolysis in the Spanish elite athlete.
in aging women: implications for osteoporosis American Journal of Sports Medicine
and fractures. Science 145:156–157. 28:57–62.
Smith S, Hayes MG, Cabana GS, et al. 2009. Sowers MR, Karvonen-Gutierrez CA. 2010. The
Inferring population continuity versus evolving role of obesity in knee osteoarthritis.
replacement with aDNA: a cautionary tale Current Opinion in Rheumatology 22:533–537.
from the Aleutian Islands. Human Biology Sparacello VS, Pearson OM, Coppa A, Marchi D.
81:19–38. 2011. Changes in skeletal robusticity in an Iron
Smrčka V. 2005. Trace Elements in Bone Tissue. Age agropastoral group: the Samnites from the
Prague, Czech Republic: Charles University. Alfedena necropolis (Abruzzo, central Italy).
Snow CC, Fitzpatrick J. 1989. Human osteological American Journal of Physical Anthropology
remains from the Battle of the Little Bighorn. 144:119–130.
In: Scott EE, Fox RAJ, Connor MA, Harmon D, Spence MW. 1974a. Residential practices and the
editors. Archaeological Perspectives on the distribution of skeletal traits in Teotihuacan,
Battle of Little Bighorn. Norman, OK: Mexico. Man 9:262–273.
University of Oklahoma Press; p. 243–306. Spence MW. 1974b. The study of residential
Snow CE. 1948. Indian Knoll Skeletons of Site practices among prehistoric hunters and
Oh 2, Ohio County, Kentucky. University of gatherers. World Archaeology 5:346–357.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 565

Spence MW. 1994. Human skeletal material from Starling AP, Stock JT. 2007. Dental indicators of
Teotihuacan. In: Sempowski ML, Spence MW, health and stress in early Egyptian and Nubian
editors. Mortuary Practices and Skeletal agriculturalists: a difficult transition and
Remains from Teotihuacan. Salt Lake City, UT: gradual recovery. American Journal of Physical
University of Utah Press; p. 315–427. Anthropology 134:520–528.
Spencer MA, Demes B. 1993. Biomechanical Stawicki SP, Grossman MD, Hoey BA, Miller DL,
analysis of masticatory system configuration Reed JF. 2004. Rib fractures in the elderly: a
in Neandertals and Inuits. American Journal of marker of injury severity. American Geriatrics
Physical Anthropology 91:1–20. Society 52:805–808.
Spielmann KA. 1991. Coercion or cooperation? Staz J. 1938. Dental caries in South Africa. South
Plains–Pueblo interaction in the Protohistoric African Journal of Medical Sciences 3
period. In: Spielmann KA, editor. Farmers, (Supplement):1–63.
Hunters, and Colonists: Interaction between Stead IM, Bourke JB, Brothwell D, editors. 1986.
the Southwest and the Southern Plains. Lindow Man: The Body in the Bog. London, UK:
Tucson, AZ: University of Arizona Press; British Museum Publications.
p. 36–50. Steadman DW. 2001. Mississippians in motion?
Spielmann KA, Schoeninger MJ, Moore K. 1990. A population genetic analysis of interregional
Plains–Pueblo interdependence and human gene flow in west-central Illinois. American
diet at Pecos Pueblo, New Mexico. American Journal of Physical Anthropology 114:61–73.
Antiquity 55:745–765. Steckel RH. 1979. Slave height profiles from
Spigelman M, Donoghue HD. 2002. The study of coastwise manifests. Explorations in Economic
ancient DNA answers a palaeopathological History 16:363–380.
question. In: Roberts CA, Lewis ME, Steckel RH. 1986. A peculiar population: the
Manchester K, editors. The Past and Present nutrition, health, and mortality of American
of Leprosy: Archaeological, Historical, slaves from childhood to maturity. Journal of
Palaeopathological and Clinical Approaches. Economic History 46:721–741.
British Archaeological Reports, International Steckel RH. 1987. Growth depression and
Series, No. 1054; p. 293–296. recovery: the remarkable case of American
Spigelman M, Lemma E. 1993. The use of slaves. Annals of Human Biology 14:111–132.
polymerase chain reaction (PCR) to detect Steckel RH. 1994. Heights and health in the United
Mycobacterium tuberculosis in ancient States, 1710–1950. In: Komlos J, editor.
skeletons. International Journal of Stature, Living Standards, and Economic
Osteoarchaeology 3:137–143. Development: Essays in Anthropometric
Sreebny LM. 1982. Sugar availability, sugar History. Chicago, IL: University of Chicago
consumption and dental caries. Community Press; p. 153–170.
Dentistry and Oral Epidemiology 10:1–7. Steckel RH. 1995. Stature and the standard of
Stafford TW, Brendel K, Duhamel RC. 1988. living. Journal of Economic Literature
Radiocarbon, 13C and 15N analysis of fossil 33:1903–1940.
bone: removal of humates with XAD-2 resin. Steckel RH. 1999. Nutritional status in the colonial
Geochimica et Cosmochimica Acta American economy. William and Mary
52:2257–2267. Quarterly 56:31–52.
Stahl AB. 1989. Plant-food processing: Steckel RH. 2005. Health and nutrition in the pre-
implications for dietary quality. In: Harris DR, Industrial era: insights from a millennium of
Hillman GC, editors. Foraging and Farming: average heights in northern Europe. In: Allen
The Evolution of Plant Exploitation. London, RC, Bengstsson T, Dribe M, editors. Living
UK: Unwin Hyman; p. 171–194. Standards in the Past: New Perspectives on
Starbuck DR. 1993. Anatomy of a massacre. Well-Being in Asia and Europe. Oxford, UK:
Archaeology 46:43–46. Oxford University Press; p. 227–253.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
566 References

Steckel RH, Floud R, editors. 1997. Health and Steffian AF, Simon JJK. 1994. Metabolic stress
Welfare during Industrialization. Chicago, IL: among prehistoric foragers of the central
University of Chicago Press. Alaskan Gulf. Arctic Anthropology 31:78–94.
Steckel RH, Rose JC, editors. 2002. The Backbone Steinbock RT. 1976. Paleopathological Diagnosis
of History: Long-Term Trends in Health and and Interpretation. Springfield, IL: Charles
Nutrition in the Americas. New York, NY: C. Thomas.
Cambridge University Press. Stenlund B. 1993. Shoulder tendinitis and
Steckel RH, Rose JC, Larsen CS, Walker PL. 2002. osteoarthrosis of the acromioclavicular joint
Skeletal health in the Western Hemisphere and their relation to sports. British Journal of
from 4000 B.C. to the present. Evolutionary Sports Medicine 27:125–130.
Anthropology 11:142–155. Steponaitis VP. 1991. Contrasting patterns of
Steegmann AT Jr. 1985. 18th century British Mississippian development. In: Earle T, editor.
military stature: growth cessation, selective Chiefdoms: Power, Economy, and Ideology.
recruiting, secular trends, nutrition at birth, cold, Cambridge, UK: Cambridge University Press;
and occupation. Human Biology 57:77–95. p. 193–228.
Steegmann AT Jr. 1986. Skeletal stature compared Steward JH. 1938. Basin-Plateau Aboriginal
to archival stature in mid-eighteenth century Sociopolitical Groups. Bureau of American
America: Ft. William Henry. American Journal Ethnology, No. 120.
of Physical Anthropology 71:431–435. Steward JH. 1940. Native cultures of the
Steegmann AT Jr, Haseley PA. 1988. Stature Intermountain (Great Basin) area. In: Kroeber
variation in the British American colonies: AL, editor. Essays in Historical Anthropology
French and Indian War records, 1755–1763. of North America, Published in Honor of John
American Journal of Physical Anthropology S. Swanton. Smithsonian Miscellaneous
75:413–421. Collections 100; p. 445–502.
Steele DG, Olive BW. 1989. Bioarcheology of the Stewart PJ, Strathern A. 2002. Violence: Theory
Region 3 Study Area. In: Hester TR, Black SL, and Ethnography. London, UK: Continuum
Steele DG, et al., editors. From the Gulf to the International Publishing Group.
Rio Grande: Human Adaptation in Central, Stewart RJC, Platt BS. 1958. Arrested growth lines
South, and Lower Pecos Texas. Arkansas in the bones of pigs on low-protein diets.
Archeological Survey Research Series, No. 33; Proceedings of the Nutrition Society 17:v–vi.
p. 93–114. Stewart TD. 1931. Dental caries in Peruvian skulls.
Steele DG, Powell JF. 1992. Peopling of America: American Journal of Physical Anthropology
paleobiological evidence. Human Biology 15:315–326
64:303–336. Stewart TD. 1949. Notas sobre esqueletos humanos
Steen SL, Lane RW. 1998. Evaluation of habitual prehistóricos hallados en Guatemala.
activities among two Alaskan Eskimo Antropología e Historia de Guatemala 1:23–24.
populations based on musculoskeletal stress Stewart TD. 1953a. Skeletal remains from Zaculeu,
markers. International Journal of Guatemala. In: Woodbury RF, Trik AS, editors.
Osteoarchaeology 8:341–353. The Ruins of Zaculeu, Guatemala. Boston, MA:
Stefan VH. 1999. Craniometric variation and United Fruit Company; p. 295–311.
homogeneity in prehistoric/protohistoric Rapa Stewart TD. 1953b. The age incidence of neural-
Nui (Easter Island) regional populations. arch defects in Alaskan natives, considered
American Journal of Physical Anthropology from the standpoint of etiology. Journal of
110:407–419. Bone and Joint Surgery 35-A:937–950.
Stefan VH, Chapman PM. 2003. Cranial Stewart TD. 1956. Examination of the possibility
variation in the Marquesas Islands. American that certain skeletal characters predispose to
Journal of Physical Anthropology defects in the lumbar neural arches. Clinical
121:319–331. Orthopaedics 8:44–60.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 567

Stewart TD. 1957. Stone Age Surgery: A General Stirland AJ. 1993. Asymmetry and activity-related
Review, with an Emphasis on the New World. change in the male humerus. International
Annual Report of the Smithsonian Institution Journal of Osteoarchaeology 3:105–113.
for 1957; p. 469–491. Stirland A. 1996. Patterns of trauma in a
Stewart TD. 1960. A physical anthropologist’s unique Medieval parish cemetery.
view of the peopling of the New World. International Journal of Osteoarchaeology
Southwestern Journal of Anthropology 6:92–100.
16:259–273. Stirland AJ. 1998. Musculoskeletal evidence for
Stewart TD. 1974. Nonunion of fractures in activity: problems of evaluation. International
antiquity, with descriptions of five cases from Journal of Osteoarchaeology 8:354–362.
the New World involving the forearm. Bulletin Stirland AJ. 2002. Raising the Dead: The Skeleton
of the New York Academy of Medicine Crew of King Henry VIII’s Great Ship, the Mary
50:875–891. Rose. Chichester, UK: John Wiley & Sons.
Stewart TD. 1977. The Neanderthal skeletal Stock JT. 2006. Hunter-gatherer postcranial
remains from Shanidar Cave, Iraq: a summary robusticity relative to patterns of mobility,
of findings to date. Proceedings of the climatic adaptation, and selection for tissue
American Philosophical Society 121:121–165. economy. American Journal of Physical
Stewart TD. 1979. Patterning of skeletal Anthropology 131:194–204.
pathologies and epidemiology. In: Laughlin Stock JT, O’Neill MC, Ruff CB, et al. 2011. Body
WS, Harper AB, editors. The First Americans: size, skeletal biomechanics, mobility and
Origins, Affinities, and Adaptations. New York, habitual activity from the Late Palaeolithic to
NY: Gustav Fischer; p. 257–274. the Mid-Dynastic Nile Valley. In: Pinhasi R,
Stewart TD, Spoehr A. 1952. Evidence on the Stock JT, editors. Human Bioarchaeology of the
paleopathology of yaws. Bulletin of the History Transition to Agriculture. Chichester, UK: John
of Medicine 26:538–553. Wiley & Sons; p. 347–367.
Steyn M, Nienaber WC. 2005. Repatriation of Stock J, Pfeiffer S. 2001. Linking structural
human remains in South Africa. In: Strkalj G, variability in long bone diaphyses to habitual
Pather N, Kramer B, editors. Voyages in behaviors: foragers from the southern Africa
Science: Essays by South African Anatomists Later Stone Age and the Andaman Islands.
in Honour of Phillip V. Tobias’s 80th Birthday. American Journal of Physical Anthropology
Pretoria, South Africa: Content Solutions; 115:337–348.
p. 159–178. Stock JT, Pfeiffer S. 2004. Long bone robusticity
Stimson BS. 1943. A Manual of Fractures and and subsistence behavior among Late Stone
Dislocations. Philadelphia, PA: Lea & Febiger. Age foragers of the forest and fynbos biomes of
Stini WA. 1990. “Osteoporosis”: etiologies, South Africa. Journal of Archaeological Science
prevention and treatment. Yearbook of 31:999–1013.
Physical Anthropology 33:151–194. Stock JT, Shaw CN. 2007. Which measures of
Stini WA. 1995. Osteoporosis in biocultural diaphyseal robusticity are robust?
perspective. Annual Review of Anthropology A comparison of external methods of
24:397–421. quantifying the strength of long bone
Stinson S. 1992. Nutritional adaptation. Annual diaphyses to cross-sectional geometric
Review of Anthropology 21:143–170. properties. American Journal of Physical
Stinson S. 2000. Growth variation: biological and Anthropology 134:412–423.
cultural factors. In: Stinson S, Bogin B, Huss- Stockmann C, Fandrey J. 2006. Hypoxia-induced
Ashmore R, O’Rourke D, editors. Human erythropoietin production: a paradigm for
Biology: An Evolutionary and Biocultural oxygen-regulated gene expression. Clinical,
Perspective. New York, NY: Wiley-Liss; Experimental Pharmacological Physiology
p. 425–463. 33:968–979.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
568 References

Stodder ALW. 1990. Paleoepidemiology of Eastern treponematosis in pre- and proto-historic


and Western Pueblo Communities in villages in western Micronesia. American
Protohistoric New Mexico. PhD Dissertation, Journal of Physical Anthropology Supplement
University of Colorado, Boulder, CO. 14:157.
Stodder ALW. 1994. Bioarchaeological Stojanowski CM. 2004. Population history of
investigations of protohistoric Pueblo health native groups in pre- and postcontact Spanish
and demography. In: Larsen CS, Milner GR, Florida: aggregation, gene flow, and genetic
editors. In the Wake of Contact: Biological drift on the southeastern U.S. Atlantic coast.
Responses to Conquest. New York, NY: Wiley- American Journal of Physical Anthropology
Liss; p. 97–107. 123:316–332.
Stodder ALW. 1997. Subadult stress, morbidity, Stojanowski CM. 2005a. The bioarchaeology of
and longevity in Latte Period populations on identity in Spanish colonial Florida: social and
Guam, Mariana Islands. American Journal of evolutionary transformation before, during,
Physical Anthropology 104:363–380. and after demographic collapse. American
Stodder ALW. 2006. Skeletal biology: Southwest. Anthropologist 107:417–431.
In: Ubelaker DH, editor. Handbook of North Stojanowski CM. 2005b. Spanish colonial effects
American Indians: 3. Environment, Origins, on Native American mating structure and
and Population. Washington, DC: Smithsonian genetic variability in northern and central
Institution Press; p. 557–580. Florida: evidence from Apalachee and western
Stodder ALW, editor. 2008. Reanalysis and Timucua. American Journal of Physical
Reinterpretation in Southwestern Anthropology 128:273–286.
Bioarchaeology. Arizona State University, Stojanowski CM. 2005c. Biocultural Histories in
Anthropological Research Papers, No. 59. La Florida. Tuscaloosa, AL: University of
Stodder ALW, Martin DL. 1992. Health and disease Alabama Press.
in the Southwest before and after Spanish Stojanowski CM. 2005d. Biological structure of
contact. In: Verano JW, Ubelaker DH, editors. the San Pedro y San Pablo de Patale Mission
Disease and Demography in the Americas. cemetery. Southeastern Archaeology
Washington, DC: Smithsonian Institution 24:165–179.
Press; p. 55–73. Stojanowski CM. 2009. Bridging histories: the
Stodder ALW, Martin DL, Goodman AH, Reff DT. bioarchaeology of identity in postcontact
2002. Cultural longevity and biological stress Florida. In: Knudson KJ, Stojanowski CM,
in the American Southwest. In: Steckel RH, editors. Bioarchaeology and Identity in the
Rose JC, editors. The Backbone of History: Americas. Gainesville, FL: University Press of
Health and Nutrition in the Western Florida; p. 59–81.
Hemisphere. New York, NY: Cambridge Stojanowski CM. 2013a. Ethnogenetic theory and
University Press; p. 481–505. new directions in biodistance research. In:
Stodder ALW, Osterholtz AJ, Mowrer K, Chuipka Lozada MC, O’Donnabhain B. The Dead Tell
JP. 2010. Processed human remains from the Tales: Essays in Honor of Jane E. Buikstra. Los
Sacred Ridge site: context, taphonomy, Angeles, CA: UCLA Cotsen Institute of
interpretation. In: Perry EM, Stodder ALW, Archaeology Press; p. 71–82.
Bollong CA, editors. Animas–La Plata Project: Stojanowski CM. 2013b. Mission Cemeteries,
Bioarchaeology. Phoenix, AZ: SWCA Mission Peoples: Historical and Evolutionary
Environmental Consultants; p. 279–415. Dimensions of Intracemetery Bioarchaeology in
Stodder ALW, Palkovich AM, editors. 2012. The Spanish Florida. Gainesville, FL: University
Bioarchaeology of Individuals. Gainesville, FL: Press of Florida.
University Press of Florida. Stojanowski CM, Buikstra JE. 2005. Research
Stodder ALW, Trembly DL, Tucker CE. 1992. trends in human osteology: a content analysis
Paleoepidemiology and paleopathology of of papers published in the American Journal of

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 569

Physical Anthropology. American Journal of Stone AC, Stoneking M. 1993. Ancient DNA from
Physical Anthropology 128:98–109. a pre-Columbian Amerindian population.
Stojanowski CM, Carver CL. 2011. Inference of American Journal of Physical Anthropology
emergent cattle pastoralism in the southern 92:463–471.
Sahara desert based on localized hypoplasia of Stone AC, Wilbur AK, Buikstra JE, Roberts CA.
the primary canine. International Journal of 2009. Tuberculosis and leprosy in perspective.
Paleopathology 1:89–97. Yearbook of Physical Anthropology 52:66–94.
Stojanowski CM, Duncan WN. 2014. Engaging Stone JL, Miles ML. 1990. Skull trepanation
bodies in the public imagination: among the early Indians of Canada and the
bioarchaeology as social science, science, and United States. Neurosurgery 26:1015–1019.
humanities. American Journal of Human Storey R. 1992a. Life and Death in the Ancient City
Biology 27:51–60. of Teotihuacan: A Modern Paleodemographic
Stojanowski CM, Johnson KM, Duncan WN. 2013. Synthesis. Tuscaloosa, AL: University of
Sinodonty and beyond: hemispheric, regional, Alabama Press.
and intracemetery approaches to studying Storey R. 1992b. Patterns of susceptibility to
dental morphological variation in the New dental defects in the deciduous dentition of a
World. In: Scott GR, Irish JD, editors. Precolumbian skeletal population. In:
Anthropological Perspectives on Tooth Goodman AH, Capasso LL, editors. Recent
Morphology: Genetics, Evolution, Variation. Contributions to the Study of Enamel
Cambridge, UK: Cambridge University Press; Developmental Defects. Journal of
p. 408–452. Paleopathology, Monographic Publications,
Stojanowski CM, Knudson KJ. 2011. No. 2; p. 61–77.
Biogeochemical inferences of mobility of early Storey R. 1992c. The children of Copán: issues in
Holocene fisher-foragers from the southern paleopathology and paleodemography.
Sahara Desert. American Journal of Physical Ancient Mesoamerica 3:161–167.
Anthropology 146:49–61. Storey R. 1998. The mothers and daughters of a
Stojanowski CM, Larsen CS, Tung TA, McEwan patrilineal civilization: the health of females
BG. 2007. Biological structure and health among Late Classic Maya of Copán. In: Grauer
implications from tooth size at Mission San AL, Stuart-Macadam P, editors. Sex and Gender
Luis de Apalachee. American Journal of in Paleopathological Perspective. Cambridge,
Physical Anthropology 132:207–222. UK: Cambridge University Press; p. 133–148.
Stojanowski CM, Schillaci MA. 2006. Phenotypic Storey R. 2005. Health and lifestyle (before and
approaches for understanding patterns of after death) among the Copán elite. In:
intracemetery biological variation. Yearbook of Andrews EW, Fash WL, editors. Copán: The
Physical Anthropology 49:49–88. History of an Ancient Maya Kingdom. Santa Fe,
Stojanowski CM, Seidemann RM, Doran GH. 2002. NM: School of American Research; p. 315–343.
Differential skeletal preservation at Windover Stout SD. 1978. Histological structures and its
Pond: causes and consequences. American preservation in ancient bone. Current
Journal of Physical Anthropology 119:15–26. Anthropology 19:601–603.
Stone AC. 2000. Ancient DNA from skeletal Stout SD. 1982. The effects of long-term
remains. In: Katzenberg MA, Saunders SR, immobilization on the histomorphology of
editors. Biological Anthropology of the Human human cortical bone. Calcified Tissue
Skeleton. New York, NY: Wiley-Liss; p. 351–371. International 34:337–342.
Stone AC. 2008. DNA analysis of archaeological Stout SD. 1983. The application of
remains. In: Katzenberg MA, Saunders SR, histomorphometric analysis to ancient skeletal
editors. Biological Anthropology of the Human remains. Anthropos 10:60–71.
Skeleton, Second Edition. Hoboken, NJ: Wiley- Stout SD. 1989. Histomorphometric analysis of
Liss; p. 461–483. human skeletal remains. In: İşcan MY,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
570 References

Kennedy KAR, editors. Reconstruction of Life Stuart-Macadam P. 1991. Anaemia in Roman


from the Skeleton. New York: Wiley-Liss; Britain: Poundbury Camp. In: Bush H,
p. 41–52. Zvelebil M, editors. Health in Past Societies:
Stout SD, Lueck R. 1995. Bone remodeling rates Biocultural Interpretations of Human Skeletal
and skeletal maturation in three archaeological Remains in Archaeological Contexts. British
skeletal populations. American Journal of Archaeological Reports, International Series,
Physical Anthropology 98:161–171. No. 567; p. 101–113.
Stout SD, Simmons DJ. 1979. Use of histology in Stuart-Macadam PL. 1989. Nutritional deficiency
ancient bone research. Yearbook of Physical diseases: a survey of scurvy, rickets, and iron-
Anthropology 22:228–249. deficiency anemia. In: İşcan MY, Kennedy
Stout SD, Teitelbaum SL. 1976. KAR, editors. Reconstruction of Life from the
Histomorphometric determination of formation Skeleton. New York, NY: Alan R. Liss;
rates of archaeological bone. Calcified Tissue p. 201–222.
Research 21:163–169. Sturmer T, Gunther K-P, Brenner H. 2000. Obesity,
Straus WL Jr, Cave AJE. 1957. Pathology and the overweight and patterns of osteoarthritis: the
posture of Neanderthal man. Quarterly Review Ulm Osteoarthritis Study. Journal of Clinical
of Biology 32:348–363. Epidemiology 53:307–313.
Strauss A, Da-Gloria P, De-Oliveira R, et al. 2013. Stynder DD, Ackermann RR, Sealy JC. 2007.
The oldest case of decapitation in the New Craniofacial variation and population
World. Poster presented in the Paleoamerican continuity during the South African Holocene.
Odyssey conference, Santa Fe, NM. American Journal of Physical Anthropology
Streeter M, Stout S, Trinkaus E, Burr D. 2010. Bone 134:489–500.
remodeling rates in Pleistocene humans are not Suarez BK. 1974. Neanderthal dental asymmetry
slower than the rates observed in modern and the probable mutation effect. American
populations: a reexamination of Abbott et al. Journal of Physical Anthropology 41:411–416.
(1996). American Journal of Physical Suckling G. 1989. Developmental defects of
Anthropology 141:315–318. enamel – historical and present-day
Strouhal E. 1971. Anthropometric and functional perspectives of their pathogenesis. Advances in
evidence of heterosis from Egyptian Nubia. Dental Research 3:87–94.
Human Biology 43:271–287. Suckling G, Elliot DC, Thurley DC. 1983. The
Strouhal E. 1991. Vertebral tuberculosis in ancient production of developmental defects of the
Egypt and Nubia. In: Ortner DJ, Aufderheide enamel in the incisor teeth of penned sheep
AC, editors. Human Paleopathology: Current resulting from induced parasitism. Archives of
Syntheses and Future Options. Washington, Oral Biology 28:393–399.
DC: Smithsonian Institution Press; Suckling GW, Elliot DC, Thurley DC. 1986. The
p. 181–194. macroscopic appearance and associated
Strouhal E, Horackova L, Likovsky J, Vargova L, histological changes in the enamel organ of
Danes J. 2002. Traces of leprosy from the Czech hypoplastic lesions of sheep incisor teeth
kingdom. In: Roberts CA, Lewis ME, resulting from induced parasitism. Archives of
Manchester K, editors. The Past and Present of Oral Biology 31:427–439.
Leprosy: Archaeological, Historical, Suckling G, Thurley DC. 1984. Developmental
Palaeopathological and Clinical Approaches. defects of enamel: factors influencing their
British Archaeological Reports, International macroscopic appearance. In: Fearnhead FW,
Series, No. 1054; p. 223–232. Suga S, editors. Tooth Enamel IV. Amsterdam,
Stuart-Macadam P. 1985. Porotic hyperostosis: the Netherlands: Elsevier; p. 357–362.
representative of a childhood condition. Sugiyama S. 1989. Burials dedicated to the Old
American Journal of Physical Anthropology Temple of Quetzalcoatl at Teotihuacan,
66:391–398. Mexico. American Antiquity 54:85–106.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 571

Sullivan A. 2004. Reconstructing relationships Sutter RC, Verano J. 2007. Biodistance analysis of
among mortality, status, and gender at the the Moche sacrificial victims from Huaca de la
Medieval Gilbertine priory of St. Andrew, Luna Plaza 3C: matrix method test of their
Fishergate, York. American Journal of Physical origins. American Journal of Physical
Anthropology 124:330–345. Anthropology 132:193–206.
Sullivan A. 2005. Prevalence and etiology of Suzuki H. 1969. Microevolutional changes in the
acquired anemia in Medieval York, England. Japanese population from the prehistoric age
American Journal of Physical Anthropology to the present-day. Journal of the Faculty of
128:252–272. Science, University of Tokyo, Section 5,
Sullivan LR. 1922. The Frequency and Distribution 3:279–308.
of Some Anatomical Variations in American Suzuki H. 1981. Racial history of the Japanese. In:
Crania. Anthropological Papers of the Schwidetzky I, editor. Rassengeschichte der
American Museum of Natural History, Menschheit. Munich, Germany: R. Oldenburg
23(part 5). Verlag; p. 7–69.
Sullivan NC. 1990. The biological consequences of Suzuki H, Yajima K, Yamanobe T, editors. 1967.
the Mississippian expansion into the western Studies on the Graves, Coffin Contents and
Great Lakes region: a study of prehistoric Skeletal Remains of the Tokugawa Shoguns and
culture contact. In: Nelson SM, Kehoe AB, Their Families at the Zojoji Temple. Tokyo,
editors. Powers of Observation: Alternative Japan: Tokyo University Press.
Views in Archeology. Archeological Papers of Suzuki T. 1982–1984. Paleopathological study on
the American Anthropological Association, osseous syphilis in skulls of the Ainu skeletal
No. 2; p. 73–87. remains. Ossa 9–11:153–168.
Sumner DR. 1985. A probable case of prehistoric Suzuki T. 1991. Paleopathological study on
tuberculosis from northeastern Arizona. In: infectious diseases in Japan. In: Ortner DJ,
Merbs CF, Miller RJ, editors. Health and Aufderheide AC, editors. Human
Disease in the Prehistoric Southwest. Arizona Paleopathology: Current Syntheses and Future
State University Anthropological Research Options. Washington, DC: Smithsonian
Papers, No. 34; p. 340–346. Institution Press; p. 128–138.
Sumner DR, Mockbee B, Morse K, Cram T, Pitt M. Suzuki T. 2013. Tuberculosis and population
1985. Computed tomography and automated movement across the Sea of Japan from the
image analysis of prehistoric femora. American Neolithic period to the Eneolithic. In:
Journal of Physical Anthropology 68:225–232. Pechenkina K, Oxenham M, editors.
Sundick RI. 1978. Human skeletal growth and age Bioarchaeology of East Asia: Movement,
determination. Homo 29:228–249. Contact, Health. Gainesville, FL: University of
Sutter RC. 1995. Dental pathologies among Florida Press; p. 125–143.
inmates of the Monroe County Poorhouse. In: Suzuki T, Fujita H, Choi JG. 2008. New evidence of
Grauer AL, editor. Bodies of Evidence: tuberculosis from prehistoric Korea –
Reconstructing History through Skeletal population movement and early evidence of
Analysis. New York, NY: Wiley-Liss; tuberculosis in Far East Asia. American Journal
p. 185–196. of Physical Anthropology 136:357–360.
Sutter RC, Cortez RJ. 2005. The nature of Moche Suzuki T, Inoue T. 2007. Earliest evidence of spinal
human sacrifice: a bio-archaeological tuberculosis from the Aneolithic Yayoi period
perspective. Current Anthropology 46: in Japan. International Journal of
521–549. Osteoarchaeology 17:392–402.
Sutter RC, Mertz L. 2004. Nonmetric cranial trait Svyatko SV, Schulting RJ, Mallory J, et al. 2013.
variation and prehistoric biocultural change in Stable isotope dietary analysis of prehistoric
the Azapa Valley, Chile. American Journal of populations from the Minusinsk Basin,
Physical Anthropology 123:130–145. Southern Siberia, Russia: a new chronological

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
572 References

framework for the introduction of millet to the study of adult Finnish women. American
eastern Eurasian steppe. Journal of Journal of Physical Anthropology 41:285–294.
Archaeological Science 40:3936–3945. Tamm E, Kivisild T, Reida M, et al. 2007. Beringian
Swanton JR. 1942. Source Material on the History standstill and spread of Native American
and Ethnology of the Caddo Indians. Bureau of founders. PLoS ONE 2:e829.
American Ethnology Bulletin, No. 132. Tanner ACR, Mathney JMJ, Kent RL Jr, et al. 2012.
Swanton JR. 1946. Indians of the Southeastern Cultivable anaerobic microbiota of severe early
United States. Bureau of American Ethnology childhood caries. Journal of Clinical
Bulletin, No. 137. Microbiology 49:1464–1474.
Swärdstedt T. 1966. Odontological Aspects of a Tanner JM, Hayashi T, Preece MA, Cameron N.
Medieval Population in the Province of 1982. Increase in length of leg relative to trunk
Jämtland/Mid-Sweden. Stockholm, Sweden: in Japanese children and adults from 1957 to
Tiden-Barnängen Tryckerier. 1977: comparisons with British and with
Swedlund AC, Armelagos GJ. 1976. Demographic Japanese Americans. Annals of Human Biology
Anthropology. Dubuque, IA: WC Brown. 9:411–423.
Sweeney EA, Saffir JA, de Leon R. 1971. Linear Tatarek NE. 2006. Geographical height variation
enamel hypoplasias of deciduous incisor teeth among Ohio Caucasian male convicts born
in malnourished children. American Journal of 1780–1849. Economics and Human Biology
Clinical Nutrition 24:29–31. 4:222–236.
Sylvester AD, Garofalo E, Ruff CB. 2010. An Tatarek NE, Sciulli PW. 2000. Comparison of
R program for automating bone cross section population structure in Ohio’s Late Archaic and
reconstruction. American Journal of Physical Late Prehistoric periods. American Journal of
Anthropology 142:665–669. Physical Anthropology 112:363–376.
Szelekovszky M, Marcsik A. 2010. Tauber H. 1981. 13C evidence for dietary habits of
Anthropological analysis of human skeletal prehistoric man in Denmark. Nature
material (7–9th century AD), east Hungary. 292:332–333.
Annuaire Roumain d’Anthropologie Tauber H. 1986. Analysis of stable isotopes in
47:3–15. prehistoric populations. In: Herrmann B, editor.
Szpunar CB, Lambert JB, Buikstra JE. 1978. Innovative Trends in Prehistoric Archaeology.
Analysis of excavated bone by atomic Berlin, Mitteilungen der Berliner Gesellschaft
absorption. American Journal of Physical für Anthropologie, Ethnologie und
Anthropology 48:199–202. Urgeschichte, 7; p. 31–38.
Tafuri MA, Craig OE, Canci A. 2009. Stable isotope Tayles N. 1999. The People. Volume 5. The
evidence for the consumption of millet and Excavation of Khok Phanom Di: A Prehistoric
other plants. American Journal of Physical Site in Central Thailand. London, UK: Society
Anthropology 139:146–153. of Antiquities of London.
Tainter JA. 1980. Behavior and status in a Middle Tayles N, Buckley HR. 2004. Leprosy and
Woodland mortuary population from the tuberculosis in Iron Age Southeast Asia?
Illinois Valley. American Antiquity American Journal of Physical Anthropology
45:308–313. 125:239–256.
Tal H. 1985. The clinical severity of periodontal Tayles N, Domett K, Halcrow S. 2009. Can dental
bone loss in dry mandibles of South African caries be interpreted as evidence of farming?
Blacks. International Journal of Skeletal The Asian perspective. Frontiers of Oral Biology
Research 12:203–210. 13:162–166.
Talbot ES. 1898. Degeneracy: Its Causes, Signs, Tayles N, Domett K, Nelsen K. 2000. Agriculture
and Results. London, UK: Walter Scott, Ltd. and dental caries? The case of rice in
Tallgren A. 1974. Neurocranial morphology and prehistoric Southeast Asia. World Archaeology
ageing – a longitudinal roentgen cephalometric 32:68–83.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 573

Tayles N, Halcrow S, Domett K. 2007. The people Gainesville, FL: University Press of Florida;
of Noen U-Loke. In: Higham CFW, Kijingam A, p. 82–112.
Talbot S, editors. The Origins of Civilization of Teaford MF, Lytle JD. 1996. Diet-induced changes
Angkor: Volume II. The Excavation of Noen in rates of human tooth microwear: a case
U-Loke Muang Kao. Bangkok: Fine Arts study involving stone-ground maize. American
Department of Thailand, p. 243–304. Journal of Physical Anthropology
Taylor GM, Crossey M, Saldanha J, Waldron T. 100:143–147.
1996. DNA from Mycobacterium tuberculosis Teaford MF, Walker A. 1984. Quantitative
identified in Mediaeval human skeletal remains differences in dental microwear between
using polymerase chain reaction. Journal of primate species with different diets, and a
Archaeological Science 23:789–798. comment on the presumed diet of Sivapithecus.
Taylor GM, Tucker K, Butler R, et al. 2013. American Journal of Physical Anthropology
Detection and strain typing of ancient 64:191–200.
Mycobacterium leprae from a medieval leprosy Temple DH. 2007. Dietary variation and stress
hospital. PLoS ONE 8(4):e62406. among prehistoric Jomon foragers from Japan.
Taylor GM, Widdison S, Brown IN, Young D, American Journal of Physical Anthropology
Molleson T. 2000. A Mediaeval case of 133:1035–1046.
lepromatous leprosy from 13–14th century Temple DH. 2008. What can variation in stature
Orkney, Scotland. Journal of Archaeological reveal about environmental differences
Science 27:1133–1138. between prehistoric Jomon foragers?
Taylor GM, Young DB, Mays SA. 2005. Genotypic Understanding the impact of systemic stress on
analysis of the earliest known prehistoric case developmental stability. American Journal of
of tuberculosis in Britain. Journal of Clinical Human Biology 20:431–439.
Microbiology 43:2236–2240. Temple DH. 2010. Patterns of systemic stress
Taylor JH. 2014. The collection of Egyptian during the agricultural transition in prehistoric
mummies in the British Museum. In: Fletcher Japan. American Journal of Physical
A, Antoine D, Hill AD, editors. Regarding the Anthropology 142:112–124.
Dead: Human Remains in the British Museum. Temple DH. 2011. Variability in dental caries
London, UK: British Museum; p. 103–114. prevalence between male and female foragers
Taylor RMS. 1963. Cause and effect of wear of from the Late/Final Jomon period: implications
teeth: further non-metrical studies of the teeth for dietary behavior and reproductive ecology.
and palate in Moriori and Maori skulls. Acta American Journal of Human Biology
Anatomica 53:97–157. 23:107–117.
Taylor RMS. 1986a. Dentistry aids anthropology. Temple DH. 2014. Plasticity and constraint in
New Zealand Dental Journal 82:109–112. response to early-life stressors among Late/
Taylor RMS. 1986b. Sealskin softening by teeth – Final Jomon period foragers from Japan:
a Maori case? Journal of the Polynesian Society evidence for life history trade-offs from
95:357–369. incremental microstructures of enamel.
Teaford MF. 1991. Dental microwear: what can it American Journal of Physical Anthropology
tell us about diet and dental function? In: 155:537–545.
Kelley MA, Larsen CS, editors. Advances in Temple DH, Kusaka S, Sciulli PW. 2011. Patterns
Dental Anthropology. New York, NY: Wiley- of social identity in relation to tooth ablation
Liss; p 342–356. among prehistoric Jomon foragers from the
Teaford MF, Larsen CS, Pastor RF, Noble VE. 2001. Yoshigo site, Aichi prefecture, Japan.
Pits and scratches: microscopic evidence of International Journal of Osteoarchaeology
tooth use and masticatory behavior in La 21:323–335.
Florida. In: Larsen CS, editor. Bioarchaeology of Temple DH, Larsen CS. 2007. Dental caries
Spanish Florida: The Impact of Colonialism. prevalence as evidence for agriculture and

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
574 References

subsistence variation among the prehistoric editors. Snake Hill: An Investigation of a


Yayoi of Japan: biocultural interpretations of Military Cemetery from the War of 1812.
an economy in transition. American Journal of Toronto, ON: Dundurn Press; p. 70–159.
Physical Anthropology 134:501–512. Thompson JL, Alfonso-Durruty MP, Crandall JJ,
Temple DH, Larsen CS. 2013. Bioarchaeological editors. 2014. Tracing Childhood:
perspectives on systemic stress during the Bioarchaeological Investigations of Early Lives
agricultural transition in prehistoric Japan. In: in Antiquity. Gainesville, FL: University Press
Pechenkina E, Oxenham M, editors. of Florida.
Bioarchaeology of East Asia: Movement, Thompson VD, Turck J. 2009. Adaptive cycles of
Contact, Health. Gainesville, FL: University coastal hunter-gatherers. American Antiquity
Press of Florida; p. 344–367. 74:255–278.
Temple DH, McGroarty JN, Guatelli-Steinberg D, Thompson VD, Turck JA. 2010. Island archaeology
Nakatsukasa M, Matsumura H. 2013. and Native American economies (2500 B.C. to
A comparative study of stress episode A.D. 1700) of the Georgia coast, USA. Journal
prevalence and duration among Jomon period of Field Archaeology 35:283–297.
foragers from Hokkaido. American Journal of Thompson VD, Worth J. 2012. Dwellers by the sea:
Physical Anthropology 152:230–238. Native American coastal adaptations along the
Temple DH, Nakatsukasa M, McGroarty JN. 2012. southern coasts of eastern North America.
Reconstructing patterns of systemic stress in a Journal of Anthropological Research
Jomon period subadult using microstructures 19:51–101.
of enamel. Journal of Archaeological Science Thoms H. 1947. The role of nutrition in pelvic
39:1634–1641. variation. American Journal of Obstetrics and
Teschler-Nicola M. 2012. The early Neolithic site Gynecology 54:62–73.
Asparn/Schletz (Lower Austria): Thoms H. 1956. Pelvimetry. London, UK: Cassel
anthropological evidence of interpersonal and Company.
violence. In: Schulting R, Fibiger L, editors. Thoms H, Foote WR, Friedman I. 1939. The clinical
Sticks, Stones, & Broken Bones: Neolithic significance of pelvic variations. American
Violence in a European Perspective. Oxford, Journal of Obstetrics and Gynecology
UK: Oxford University Press; p. 101–120. 38:634–642.
Thijn CJP, Steensma JT. 1990. Tuberculosis of the Thordeman B. 1939. Armour from the Battle of
Skeleton: Focus on Radiology. Berlin, Germany: Wisby 1361. Stockholm, Sweden: Vitterhets
Springer-Verlag. Historie och Antikvitets Akademien.
Thomas DH. 1985. Why bother digging Hidden Tiesler V, Cucina A. 2006. Janaab’ Pakal of
Cave again? In: Thomas DH, editor. The Palenque: Reconstructing the Life and Death of
Archaeology of Hidden Cave, Nevada. a Maya Ruler. Tucson, AZ: University of
Anthropological Papers of the American Arizona Press.
Museum of Natural History, 61(part 1):17–20. Tiesler V, Cucina A, editors. 2007. New
Thomas DH. 1988. The Archaeology of Monitor Perspectives on Human Sacrifice and Ritual
Valley, part I: Epistemology. Anthropological Body Treatments in Ancient Maya Society. New
Papers of the American Museum of Natural York, NY: Springer.
History, 58. Tieszen LL. 1991. Natural variations in the carbon
Thomas DH, South S, Larsen CS. 1977. Rich Man, isotope values of plants: implications for
Poor Men: Observations on Three Antebellum archaeology, ecology and paleoecology.
Burials from the Georgia Coast. Journal of Archaeological Science 18:227–248.
Anthropological Papers of the American Tieszen LL, Boutton TW. 1989. Stable carbon
Museum of Natural History, 54(part 3). isotopes in terrestrial ecosystem research. In:
Thomas SC, Williamson RF. 1991. Archaeological Baillie TA, Jones JR, editors. Synthesis and
investigations. In: Pfeiffer S, Williamson RF, Applications of Isotopically Labeled

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 575

Compounds. Amsterdam, the Netherlands: and elbows of preadolescent baseball players.


Elsevier; p. 13–19. Pediatrics 49:267–272.
Tieszen LL, Boutton TW, Tesdahl KG, Slade NA. Torres-Rouff C. 2011. Hiding inequality beneath
1983. Fractionation and turnover of stable prosperity: patterns of cranial injury in Middle
carbon isotopes in animal tissues: implications Period San Pedro de Atacama, northern Chile.
for the δ13C analysis of diet. Oecologia American Journal of Physical Anthropology
57:32–37. 146:28–37.
Tieszen LL, Fagre T. 1993. Effects of diet quality Torres-Rouff C, Costa Junqueira MA. 2006.
and composition on the isotopic composition Interpersonal violence in prehistoric San Pedro
of respiratory CO2, bone collagen, bioapatite de Atacama, Chile: behavioral implications of
and soft tissues. In: Lambert JB, Grupe G, environmental stress. American Journal of
editors. Prehistoric Human Bone: Archaeology Physical Anthropology 130:60–70.
at the Molecular Level. Berlin: Springer-Verlag; Torres-Rouff C, Knudson KJ, Hubbe M. 2013.
p. 121–155. Issues of affinity: exploring population
Timoshenko SP, Gere JM. 1972. Mechanics of structure in the Middle and Regional
Materials. New York, NY: Van Nostrand Development periods of San Pedro de Atacama,
Reinhold. Chile. American Journal of Physical
Tobias PV. 1967. The Cranium and Maxillary Anthropology 152:370–382.
Dentition of Zinjanthropus (Australopithecus) Torroni A, Schurr TG, Campbell MF, et al. 1993.
boisei. Cambridge, UK: Cambridge University Asian affinities and continental radiation of
Press. the four founding Native American mtDNAs.
Todd TW. 1920. Age changes in the pubic bone: American Journal of Human Genetics
I. The male White pubis. American Journal of 53:563–590.
Physical Anthropology 3:285–334. Tóth K. 1970. The Epidemiology of Dental Caries in
Todd TW. 1924. Thickness of the male Hungary. Budapest: Akadémiai Kiadó.
White cranium. Anatomical Record Toth N. 1985. Archaeological evidence for
27:245–256. preferential right-handedness in the Lower and
Todd TW. 1927. The immediate appearance of a Middle Pleistocene, and its possible implications.
fracture of the lower extremity of the radius. Journal of Human Evolution 14:607–614.
Annals of Surgery 81:956–958. Toverud G, Cox GJ, Finn SB, Bodecker CF. 1952.
Todd TW, Barber CG. 1934. The extent of skeletal A Survey of the Literature of Dental Caries.
change after amputation. Journal of Bone and National Academy of Sciences, National
Joint Surgery 16-A:53–64. Research Council Publication, 225.
Tomczak PD. 2003. Prehistoric diet and Townsend G, Dempsey P, Brown T, Kaidonis J,
socioeconomic relationships within the Richards L. 1994. Teeth, genes and the
Osmore Valley of southern Peru. Journal of environment. Perspectives in Human Biology
Anthropological Archaeology 22:262–278. 4:35–46.
Tomczak PD, Powell JF. 2003. Postmarital Townsend G, Farmer V. 1998. Dental asymmetry
residence practices in the Windover in the deciduous dentition of South Australian
population: sex-based dental variation as an children. In: Lukacs JR, editor. Human Dental
indicator of patrilocality. American Antiquity Development, Morphology, and Pathology:
68:93–108. A Tribute to Albert A. Dahlberg. University of
Tomenchuk J, Mayhall JT. 1979. A correlation of Oregon Anthropological Papers, No. 54;
tooth wear and age among modern Igloolik p. 245–285.
Eskimos. American Journal of Physical Townsend G, Richards L, Hughes T. 2003. Molar
Anthropology 51:67–78. intercuspal dimensions: genetic input to
Torg JS, Pollack H, Sweterlitsch P. 1972. The phenotypic variations. Journal of Dental
effects of competitive pitching on the shoulders Research 82:350–355.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
576 References

Townsend GC. 1980. Heritability of deciduous Trinkaus E. 1993. Femoral neck–shaft angles of
tooth size in Australian Aboriginals. American the Qafzeh-Skhul early modern humans, and
Journal of Physical Anthropology 53:297–300. activity levels among immature Near Eastern
Townsend GC. 1992. Genetic and environmental Middle Paleolithic hominids. Journal of Human
contributions to morphometric dental Evolution 25:393–416.
variation. In: Lukacs JR, editor. Culture, Trinkaus E, Churchill SE, Ruff CB. 1994.
Ecology and Dental Anthropology. Journal of Postcranial robusticity in Homo: II. Humeral
Human Ecology Special Issue, 2: 61–72. bilateral asymmetry and bone plasticity.
Townsend GC, Brown T. 1978. Heritabilities of American Journal of Physical Anthropology
permanent tooth size. American Journal of 93:1–34.
Physical Anthropology 49:497–502. Trinkaus E, Ruff CB. 1989. Cross-sectional
Townsend GC, Brown T. 1980. Dental asymmetry geometry of Neandertal femoral and tibial
in Australian Aborigines. Human Biology diaphyses: implications for locomotion.
52:661–673. American Journal of Physical Anthropology
Townsley W. 1946. Platymeria. Journal of 78:315–316.
Pathology and Bacteriology 58:85–88. Trinkaus E, Ruff CB. 1999. Diaphyseal cross-
Tran T-N-N, Le Forestier C, Drancourt M, Raoult D, sectional geometry of Near Eastern Middle
Aboudharam G. 2011. Co-detection of Palaeolithic humans: the tibia. Journal of
Bartonella quintana and Yersinia pestis in an Archaeological Science 26:1289–1300.
11th–15th-century burial site in Bondy, Trinkaus E, Ruff CB. 2012. Femoral and tibial
France. American Journal of Physical diaphyseal cross-sectional geometry in
Anthropology 145:489–494. Pleistocene Homo. PaleoAnthropology
Trembly D. 1995. Spondylolysis and 2012:13–62.
spondylolisthesis. Paleopathology Newsletter Trinkaus E, Zimmerman MR. 1982. Trauma among
92:14. the Shanidar Neandertals. American Journal of
Trembly DL. 2002. Perspectives on the history of Physical Anthropology 57:61–76.
leprosy in the Pacific. In: Roberts CA, Lewis Tucker K. 2014. The osteology of decapitation
ME, Manchester K, editors. The Past and burials from Roman Britain: a post-mortem
Present of Leprosy: Archaeological, Historical, burial rite? In: Knüsel C, Smith MJ, editors. The
Palaeopathological and Clinical Approaches. Routledge Handbook of the Bioarchaeology of
British Archaeological Reports, International Human Conflict. London, UK: Routledge;
Series, No. 1054; p. 233–238. p. 213–236.
Trinkaus E. 1975. Squatting among the Tung TA. 2007. From corporeality to sanctity:
Neandertals: a problem in the behavioral transforming bodies into trophy heads in the
interpretation of skeletal morphology. Journal pre-Hispanic Andes. In: Chacon RJ, Dye DH,
of Archaeological Science 2:327–351. editors. The Taking and Displaying of Human
Trinkaus E. 1977. The Alto Salaverry child: a case of Body Parts as Trophies by Amerindians. New
anemia from the Peruvian Preceramic. American York, NY: Springer; p. 481–504.
Journal of Physical Anthropology 46:25–28. Tung TA. 2008. Dismembering bodies for display: a
Trinkaus E. 1983. The Shanidar Neandertals. New bioarchaeological study of trophy heads from the
York, NY: Academic Press. Wari site of Conchopata, Peru. American Journal
Trinkaus E. 1984. Western Asia. In: Smith FH, of Physical Anthropology 136:294–308.
Spencer F, editors. The Origins of Modern Tung TA. 2012a. Violence, Ritual, and the Wari
Humans: A World Survey of the Fossil Evidence. Empire: A Social Bioarchaeology of
New York, NY: Alan R. Liss; p. 251–293. Imperialism in the Ancient Andes. Gainesville,
Trinkaus E. 1985. Pathology and the posture of the FL: University Press of Florida.
La Chapelle-aux-Saints Neandertal. American Tung TA. 2012b. Violence against women:
Journal of Physical Anthropology 67:19–41. differential treatment of local and foreign

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 577

females in the heartland of the Wari empire, Turner CG II. 1986a. Dentochronological
Peru. In: Martin DL, Harrod RP, Perez VR, separation estimates for Pacific rim
editors. The Bioarchaeology of Violence. populations. Science 232:1140–1142.
Gainesville, FL: University Press of Florida; Turner CG II. 1986b. The first Americans: the
p. 180–198. dental evidence. National Geographic Research
Tung TA, Knudson KJ. 2008. Social identities and 2:37–46.
geographical origins of Wari trophy heads Turner CG II. 1987. Late Pleistocene and Holocene
from Conchopata, Peru. Current Anthropology population history of East Asia based on dental
49:915–925. variation. American Journal of Physical
Tung TA, Knudson KJ. 2011. Identifying locals, Anthropology 73:305–321.
migrants, and captives in the Wari heartland: a Turner CG II. 1991. The Dentition of Arctic
bioarchaeological and biogeochemical study of Peoples. New York, NY: Garland Publishing.
human remains from Conchopata, Peru. Turner CG II. 1993a. Cannibalism in Chaco
Journal of Anthropological Archaeology Canyon: the charnel pit excavated in 1926 at
30:247–261. Small House Ruin by Frank H.H. Roberts, Jr.
Tuominen M, Kantomaa T, Pirttiniemi P. 1993. American Journal of Physical Anthropology
Effect of food consistency on the shape of the 91:421–439.
articular eminence and the mandible: an Turner CG II. 1993b. Southwest Indian teeth.
experimental study in the rabbit. Acta National Geographic Research & Exploration
Odontologica Scandinavica 51:65–72. 9:32–53.
Turner BL, Kamernov GD, Kingston JD, Armelagos Turner CG II, Bird J. 1981. Dentition of Chilean
GJ. 2009. Insights into immigration and Paleo-Indians and peopling of the Americas.
social class at Machu Picchu, Peru based on Science 212:1053–1055.
oxygen, strontium, and lead isotope Turner CH, Burr DB. 1993. Basic biomechanical
analysis. Journal of Archaeological Science measurements of bone: a tutorial. Bone
36:317–332. 14:595–608.
Turner CG. 2004. A second drilled tooth from Turner CG II, Cadien JD. 1969. Dental chipping in
prehistoric western North America. American Aleuts, Eskimos and Indians. American Journal
Antiquity 69:356–360. of Physical Anthropology 31:303–310.
Turner CG II. 1979. Dental anthropological Turner CG II, Machado LMC. 1983. A new dental
indications of agriculture among the Jomon wear pattern and evidence for high
people of central Japan: X. Peopling of the carbohydrate consumption in a Brazilian
Pacific. American Journal of Physical Archaic skeletal population. American Journal
Anthropology 51:619–636. of Physical Anthropology 61:125–130.
Turner CG II. 1983. Taphonomic reconstructions of Turner CG II, Markowitz MA. 1990. Dental
human violence and cannibalism based on discontinuity between Late Pleistocene and
mass burials in the American Southwest. In: recent Nubians: peopling of the Eurafrican-
LeMoine GM, MacEachern, editors. Carnivores, South Asian triangle I. Homo 41:32–41.
Human Scavengers and Predators: A Question Turner CG II, Morris NT. 1970. A massacre at Hopi.
of Bone Technology. Proceedings of the 15th American Antiquity 35:320–331.
Annual Conference, Archaeological Turner CG II, Nichol CR, Scott GR. 1991. Scoring
Association of the University of Calgary; procedures for key morphological traits of the
p. 219–240. permanent dentition. In: Kelley MA, Larsen CS,
Turner CG II. 1985. The dental search for Native editors. Advances in Dental Anthropology. New
American origins. In: Kirk R, Szathmáry E, York, NY: Wiley-Liss; p. 13–31.
editors. Out of Asia: Peopling of the Americas Turner CG II, Turner JA. 1992. The first claim for
and the Pacific. Journal of Pacific History; cannibalism in the Southwest: Walter Hough’s
p. 31–78. 1901 discovery at Canyon Butte Ruin 3,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
578 References

northeastern Arizona. American Antiquity Ubelaker DH. 1988. Prehistoric human biology at
57:661–682. La Tolita, Ecuador, a preliminary report.
Turner CG II, Turner JA. 1995. Cannibalism in the Journal of the Washington Academy of
prehistoric American Southwest: occurrence, Sciences 78:23–37.
taphonomy, explanation, and suggestions for Ubelaker DH. 1992a. Porotic hyperostosis in
standardized world definition. Anthropological prehistoric Ecuador. In: Stuart-Macadam P,
Science 103:1–22. Kent S, editors. Diet, Demography, and
Turner CG II, Turner JA. 1999. Man Corn: Disease: Changing Perspectives on Anemia.
Cannibalism and Violence in the Prehistoric Hawthorne, NY: Aldine de Gruyter;
American Southwest. Salt Lake City, UT: p. 201–217.
University of Utah Press. Ubelaker DH. 1992b. Enamel hypoplasia in
Tykot RH, Burger RL, van der Merwe NJ. 2006. The ancient Ecuador. In: Goodman AH, Capasso LL,
importance of maize in the Initial period and editors. Recent Contributions to the Study of
Early Horizon Peru. In: Staller J, Tykot R, Benz Enamel Developmental Defects. Journal of
B, editors. Histories of Maize. Amsterdam, the Paleopathology, Monographic Publications,
Netherlands: Elsevier; p. 187–197. No. 2; p. 207–217.
Tykot RH, Staller JE. 2002. The importance of Ubelaker DH. 1994. The biological impact of
early maize agriculture in coastal Ecuador: new European contact in Ecuador. In: Larsen CS,
data from La Emerenciana. Current Milner GR, editors. In the Wake of Contact:
Anthropology 43:666–677. Biological Responses to Conquest. New York,
Tyrrell A. 2000. Skeletal non-metric traits and the NY: Wiley-Liss; p. 147–160.
assessment of inter- and intra-population Ubelaker, DH. 1999. Human Skeletal Remains:
diversity: past problems and future potential. Excavation, Analysis, Interpretation, third
In: Cox M, Mays S, editors. Human Osteology edition. Washington, DC: Taraxacum.
in Archaeology and Forensic Science. London, Ubelaker DH. 2000. Human Remains from La
UK: Greenwich Medical Media; p. 289–306. Florida, Quito, Ecuador. Smithsonian
Tyson RA. 1977. Historical accounts as aids to Contributions to Anthropology, No. 43.
physical anthropology: examples of head Ubelaker DH. 2002. Approaches to the study of
injury in Baja California. Pacific Coast commingling in human skeletal biology. In:
Archaeological Society Quarterly 13:52–58. Haglund WD, Sorg MH, editors. Advances in
Ubelaker DH. 1974. Reconstruction of Forensic Taphonomy. Washington, DC: CRC
Demographic Profiles from Ossuary Skeletal Press; p. 331–351.
Samples: A Case Study from the Tidewater Ubelaker DH, Grant LG. 1989. Human skeletal
Potomac. Smithsonian Contributions to remains: preservation or reburial? Yearbook of
Anthropology, No. 18. Physical Anthropology 32:249–287.
Ubelaker DH. 1979. Skeletal evidence for kneeling Ubelaker DH, Jantz RL. 1979. Plains Caddoan
in prehistoric Ecuador. American Journal of relationships: the view from craniometry and
Physical Anthropology 51:679–686. mortuary analysis. In: Wedel WR, editor.
Ubelaker DH. 1981. The Ayalan Cemetery: A Late Toward Plains Caddoan Origins: A
Integration Period Burial Site on the South Symposium. Nebraska History, No. 60;
Coast of Ecuador. Smithsonian Contributions p. 249–259.
to Anthropology, No. 29. Ubelaker DH, Jantz RL. 1986. Biological History of
Ubelaker DH. 1984. Prehistoric human biology of the Aboriginal Population of North America.
Ecuador: possible temporal trends and cultural Sonderdruck aus Rassengeschichte der
correlations. In: Cohen MN, Armelagos GJ, Menschheit. Munich, Germany: R. Oldenbourg
editors. Paleopathology at the Origins of Verlag.
Agriculture. Orlando, FL: Academic Press; Ubelaker DH, Jones EB. 2003. Human Remains
p. 491–513. from Voegtly Cemetery, Pittsburgh,

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 579

Pennsylvania. Smithsonian Contributions to The Larsen Bay Repatriation and the


Anthropology, No. 46. Smithsonian Institution. Washington, DC:
Ubelaker DH, Katzenberg MA, Doyon LG. 1995. Smithsonian Institution Press; p. 101–121.
Status and diet in precontact highland Valdes AM, Spector TD. 2008. The contribution of
Ecuador. American Journal of Physical genes to osteoarthritis. Rheumatic Disease
Anthropology 97:403–411. Clinics of North America 34:581–603.
Ubelaker DH, Newson LA. 2002. Patterns of health Valentin F, Bocherens H, Gratuze B, Sand C. 2006.
and nutrition in prehistoric and historic Dietary patterns during the late prehistoric/
Ecuador. In: Steckel RH, Rose JC, editors. The historic period in Cikobia Island (Fiji): insights
Backbone of History. New York, NY: Cambridge from stable isotopes and dental pathologies.
University Press; p. 343–375. Journal of Archaeological Science
Ullinger JM, Sheridan SG, Hawkey DE, Turner 33:1396–1410.
CG II, Cooley R. 2005. Bioarchaeological Valentin F, Herrscher E, Petchey F, Addison DJ.
analysis of cultural transition in the southern 2011. An analysis of the last 1000 years of
Levant using dental nonmetric traits. American human diet on Tutuila (American Somoa) using
Journal of Physical Anthropology carbon and nitrogen stable isotope data.
128:466–476. American Antiquity 76:473–486.
Ullinger JM, Sheridan SG, Ortner DJ. 2012. Daily Vallois HV. 1960. Vital statistics in prehistoric
activity and lower limb modification at Bab population as determined from archaeological
edh-Dhra’, Jordan, in the Early Bronze Age. In: data. In: Heizer RF, Cook SF, editors. The
Perry MA, editor. Bioarchaeology and Application of Quantitative Methods in
Behavior: The People of the Ancient Near East. Archaeology. New York, NY: Viking Fund;
Gainesville, FL: University Press of Florida; p. 186–222.
p. 180–201. van den Berg W. 1999. Osteophyte formation in
Ungar PS. 1994. Incisor microwear in Sumatran osteoarthritis. Osteoarthritis and Cartilage
anthropoid primates. American Journal of 7:333.
Physical Anthropology 94:339–363. van der Merwe NJ. 1982. Carbon isotopes,
Ungar PS, Scott RS, Scott JR, Teaford M. 2008. photosynthesis, and archaeology. American
Dental microwear analysis: historical Scientist 70:596–606.
perspectives and new approaches. In: Irish JD, van der Merwe NJ. 1992. How bones reveal diet.
Nelson GC, editors. Technique and Application In: Jones S, Martin R, Pilbeam D, editors. The
in Dental Anthropology. Cambridge, UK: Cambridge Encyclopedia of Human Evolution.
Cambridge University Press; p. 389–425. Cambridge, UK: Cambridge University Press;
Ungar PS, Spencer MA. 1999. Incisor microwear, p. 391.
diet, and tooth use in three Amerindian van der Merwe NJ, Lee-Thorp JA, Raymond JS.
populations. American Journal of Physical 1993. Light, stable isotopes and the subsistence
Anthropology 109:387–396. base of formative cultures at Valdavia,
Ungar PS, Teaford MF. 1996. Preliminary Ecuador. In: Lambert JB, Grupe G, editors.
examination of non-occlusal dental microwear Prehistoric Human Bone: Archaeology at the
in anthropoids: implications for the study of Molecular Level. Berlin, Germany: Springer-
fossil primates. American Journal of Physical Verlag; p. 63–97.
Anthropology 100:101–113. van der Merwe NJ, Vogel JC. 1978. 13C content of
United Nations. 2011. World Population human collagen as a measure of prehistoric
Prospects, The 2010 Revision. Department of diet in Woodland North America. Nature
Economic and Social Affairs, Washington, DC. 276:815–816.
Urcid J. 1994. Cannibalism and curated skulls: van der Merwe NJ, Williamson RF, Pfeiffer S,
bone ritualism on Kodiak Island. In: Bray TL, Thomas SC, Allegretto KO. 2003. The Moatfield
Killion TW, editors. Reckoning with the Dead: ossuary: isotopic dietary analysis of an

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
580 References

Iroquoian community, using dental tissue. Symposium on Dental Morphology, Paris 1986.
Journal of Anthropological Archaeology Mémoires Musée National, Naturale Histoire,
22:245–261. 53(Série C); p. 415–444.
Van Doren M, editor. 1928. Travels of William Van Tilburg JA. 1995. Easter Island: Archaeology,
Bartram. New York, NY: Dover. Ecology, and Culture. Washington, DC:
Van Gerven DP. 1973. Thickness and area Smithsonian Institution Press.
measurements as parameters of skeletal Van Valen L. 1962. A study of fluctuating
involution of the humerus, femur, and tibia. asymmetry. Evolution 16:125–142.
Journal of Gerontology 28:40–45. van Winkelhoff AJ, Slots J. 1999. Actinobacillus
Van Gerven DP, Armelagos GJ. 1983. Farewell to actinomycetemcomitans and Porphyromonas
paleodemography? Rumors of its death have gingivalis in nonoral infections. Periodontology
been greatly exaggerated. Journal of Human 20:122–135.
Evolution 12:353–360. Varela HH, O’Brien TG, Cocilovo JA. 2008. The
Van Gerven DP, Armelagos GJ, Rohr A. 1977. genetic divergence of prehistoric populations
Continuity and change in cranial morphology of the south-central Andes as established by
of three Nubian archaeological populations. means of craniometric traits. American Journal
Man 12:270–277. of Physical Anthropology 137:274–282.
Van Gerven DP, Carlson DS, Armelagos GJ. 1973. Varner JG, Varner JJ, editors. 1980. The Florida of
Racial history and bio-cultural adaptation of the Inca, by Garcilaso de la Vega. Austin, TX:
Nubian archaeological populations. Journal of University of Texas Press.
African History 14:555–564. Vega Lizama EM, Cucina A. 2014. Maize
Van Gerven DP, Hummert JR, Burr DB. 1985. dependence or market integration? Caries
Cortical bone maintenance and geometry of prevalence among indigenous Maya
the tibia in prehistoric children from Nubia’s communities with maize-based versus
Batn el Hajar. American Journal of Physical globalized economies. American Journal of
Anthropology 66:275–280. Physical Anthropology 153:190–202.
Van Gerven DP, Sheridan SG, Adams WY. 1995. Verano JW. 1986. A mass burial of mutilated
The health and nutrition of a Medieval Nubian individuals at Pacatnamu. In: Donnan CB,
population: the impact of political and Cock GA, editors. The Pacatnamu Papers,
economic change. American Anthropologist Volume 1. Los Angeles, CA: Museum of
97:468–480. Cultural History, University of California;
van Reenen JF. 1964. Dentition, jaws and palate of p. 117–138.
the Kalahari Bushmen. Journal of the Dental Verano JW. 1995. Where do they rest? The
Association of South Africa 19:1–65. treatment of human offerings and trophies in
van Reenen JF. 1982. The effects of attrition on ancient Peru. In: Dillehay TD, editor. Tombs for
tooth dimensions of San (Bushmen). In: Kurtén the Living: Andean Mortuary Practices.
B, editor. Teeth: Form, Function, and Evolution. Washington, DC: Dumbarton Oaks Research
New York, NY: Columbia University Press; Library and Collection; p. 189–226.
p. 182–203. Verano JW. 2001. The physical evidence of human
van Reenen JF. 1992. Dental wear in San sacrifice in ancient Peru. In: Benson ET, Cook AG,
(Bushman). In: Lukacs JR, editor. Culture, editors. Ritual Sacrifice in Ancient Peru. Austin,
Ecology, and Dental Anthropology. Delhi, TX: University of Texas Press; p. 165–184.
India: Kamla-Raj; p. 201–213. Verano JW. 2003. Trepanation in prehistoric
van Reenen JF, Reinach SG. 1988. Interstitial and South America: geographic and temporal
occlusal wear of first molar teeth in San trends over 2,000 years. In: Arnott R, Finger S,
(Bushman). In: Russell DE, Santoro J-P, Smith CUM, editors. Trepanation: History,
Sigogneau-Russell D, editors. Teeth Revisited. Discovery, Theory. Lisse, the Netherlands:
Proceedings of the VIIth International Swets & Zeitlinger Publishers; p. 223–236.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 581

Verano JW. 2005. Human sacrifice and evidence from the upper limb. American
postmortem modification at the Pyramid of the Journal of Physical Anthropology
Moon, Moche Valley, Peru. In: Rakita GFM, 142:224–234.
Buikstra JE, Beck LA, Williams SR, editors. Villotte S, Churchill SE, Dutour O, Henry-Gambier
Interacting with the Dead: Perspectives on the D. 2010b. Subsistence activities and the sexual
Mortuary Archaeology for the New Millennium. division of labor in European Upper Paleolithic
Gainesville, FL: University Press of Florida; and Mesolithic evidence from upper limb
p. 277–289. enthesiopathies. Journal of Human Evolution
Verano JW. 2007. Conflict and conquest in pre- 59:35–43.
Hispanic Andean South America. In: Chacon Villotte S, Knüsel CJ. 2013. Understanding
RJ, Mendoza RG, editors. Latin American enthesial changes: definition and life course
Indigenous Warfare and Ritual Violence. changes. International Journal of
Tucson, AZ: University of Arizona Press; Osteoarchaeology 23:135–146.
p. 105–115. Villotte S, Knüsel CJ. 2014. “I sing of arms and of a
Verano JW. 2008. Trophy head-taking and human man. . .”: medial epicondylosis and the sexual
sacrifice in Andean South America. In: division of labour in prehistoric Europe.
Silverman H, Isbell WH, editors. Handbook of Journal of Archaeological Science 43:168–174.
South American Archaeology. New York, NY: Vitzthum VJ. 1994. Comparative study of
Springer; p. 1047–1060. breastfeeding structure and its relation to
Verano JW. 2014. Many faces of death: warfare, human reproductive ecology. Yearbook of
human sacrifice and mortuary patterns of the Physical Anthropology 37:307–349.
elite in late pre-Hispanic northern Peru. In: Vladescu M. 1992. Aspects of the microevolution
Knüsel C, Smith MJ, editors. The Routledge of the cephalic index in Roumanian
Handbook of the Bioarchaeology of Human populations. Annuaire Roumain
Conflict. London, UK: Routledge; p. 355–370. d’Anthropologie 29:29–35.
Verano JW, Anderson LS, Franco R. 2000. Foot Vogel JC, van der Merwe NJ. 1977. Isotopic
amputation by the Moche of ancient Peru: evidence for early maize cultivation in New
osteological evidence and archaeological York State. American Antiquity 42:238–242.
context. International Journal of von Cramon-Taubadel N. 2009a. Congruence of
Osteoarchaeology 10:177–188. individual cranial bone morphology and
Verano JW, Ubelaker DH, editors. 1992. Disease neutral molecular affinity patterns in modern
and Demography in the Americas. Washington, humans. American Journal of Physical
DC: Smithsonian Institution Press. Anthropology 140:205–215.
Vercellotti G, Stout SD, Boano R, Sciulli PW. 2011. von Cramon-Taubadel N. 2009b. Revisiting the
Intrapopulation variation in stature and body homoiology hypothesis: the impact of
proportions: social status and sex difference in phenotypic plasticity on the reconstruction of
an Italian medieval population (Trino human population history from craniometric
Vercellese, VC). American Journal of Physical data. Journal of Human Evolution 57:179–190.
Anthropology 145:203–214. von Cramon-Taubadel N. 2011. Global human
Vieth R. 2003. Effects of vitamin D on bone and mandibular variation reflects differences in
natural selection of skin color: how much agricultural and hunter-gatherer subsistence
vitamin D nutrition are we talking about? In: strategies. Proceedings of the National
Agarwal SC, Stout SD, editors. Bone Loss and Academy of Sciences 108:19546–19551.
Osteoporosis: An Anthropological Perspective. von Hunnius TE, Roberts CA, Boylston A,
New York, NY: Kluwer Academic Plenum; Saunders SA. 2006. Histological identification
p. 139–154. of syphilis in pre-Columbian England.
Villotte S, Castex D, Couallier V, et al. 2010a. American Journal of Physical Anthropology
Enthesopathies as occupational stress markers: 129:559–566.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
582 References

von Hunnius TE, Yang D, Eng B, Waye JS, Waldron T. 1992. Unilateral spondylolysis.
Saunders SR. 2007. Digging deeper into the International Journal of Osteoarchaeology
limits of ancient DNA research on syphilis. 2:177–181.
Journal of Archaeological Science Waldron T. 1993. The health of the adults. In:
34:2091–2100. Molleson T, Cox M, editors. The Spitalfields
Wada E. 1980. Nitrogen isotope fractionation and Project: Volume 2. The Anthropology, The
its significance in biogeochemical processes Middling Sort. Council for British Archaeology,
occurring in marine environments. In: CBA Research Report, No. 86; p. 67–89.
Goldberg ED, Horibe Y, Saruhaski K, editors. Waldron T. 1994. Counting the Dead: The
Isotope Marine Chemistry. Tokyo, Japan: Epidemiology of Skeletal Populations.
Uchida Rokakuho; p. 375–398. Chichester, UK: John Wiley & Sons.
Wadsworth GR. 1992. Physiological, pathological, Waldron T. 1995. Changes in distribution of
and dietary influences on the hemoglobin osteoarthritis over historical time. International
level. In: Stuart-Macadam P, Kent S, editors. Journal of Osteoarchaeology 5:385–389.
Diet, Demography, and Disease: Changing Waldron T. 2009. Palaeopathology. Cambridge,
Perspectives on Anemia. Hawthorne, NY: UK: Cambridge University Press.
Aldine de Gruyter; p. 63–104. Waldron T. 2012. Joint disease. In: Grauer AL,
Wahl J, Trautmann I. 2012. The Neolithic editor. A Companion to Paleopathology.
massacre at Talheim: a pivotal find in conflict Chichester, UK, Wiley-Blackwell; p. 513–530.
archaeology. In: Schulting R, Fibiger L, editors. Waldron T, Rogers J. 1991. Inter-observer
Sticks, Stones, & Broken Bones: Neolithic variation in coding osteoarthritis in human
Violence in a European Perspective. Oxford, skeletal remains. International Journal of
UK: Oxford University Press; p. 77–100. Osteoarchaeology 1:49–56.
Wainwright SA. 1988. Axis and Circumference: Walimbe SR, Gambhir PB. 1994. Long Bone
The Cylindrical Shape of Plants and Animals. Growth in Infants and Children: Assessment of
Cambridge, MA: Harvard University Press. Nutritional Status. Monograph Series on
Wakely J, Bruce MF. 1989. Interpreting signs of Biological Anthropology, II. Mangalore, India:
trauma on a human axis vertebra. Journal of Mujumdar Publications.
Anatomy 167:265. Walimbe SR, Kulkarni SS. 1993. Biological
Waldron T. 1981. Postmortem absorption of lead Adaptations in Human Dentition: An
by the skeleton. American Journal of Physical Odontometric Study on Living and
Anthropology 55:395–398. Archaeological Populations in India. Pune,
Waldron T. 1982. Human bone lead India: Deccan College Post Graduate and
concentrations. In: McWhirr A, Viner L, Wells Research Institute.
C, editors. Cirencester Excavations II: Romano- Walimbe SR, Tavares A. 2002. Human skeletal
British Cemeteries at Cirencester. Cirencester, biology: scope, development and present status
UK: Corinium Museum; p. 203–207. of research in India. In: Paddayya K, editor.
Waldron T. 1983. On the post-mortem Recent Studies in Indian Archaeology. New
accumulation of lead by skeletal tissues. Delhi, India: Munshieam Manoharolal;
Journal of Archaeological Science 10:35–40. p. 367–402.
Waldron T. 1991a. Variations in the rates of Walker A. 1980. Functional anatomy and
spondylolysis in early populations. taphonomy. In: Behrensmeyer AK, Hill AP,
International Journal of Osteoarchaeology editors. Fossils in the Making: Vertebrate
1:63–65. Taphonomy and Paleoecology. Chicago, IL:
Waldron T. 1991b. Variations in the prevalence of University of Chicago Press; p. 182–196.
spondylolysis in early British populations. Walker A. 1981. Dietary hypotheses and human
Journal of the Royal Society of Medicine evolution. Philosophical Transactions of the
84:547–549. Royal Society of London Series B 292:57–64.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 583

Walker J, MacGillivray I, Macnaughton MC, Walker PL. 2001a. A bioarchaeological perspective


editors. 1976. Combined Textbook of Obstetrics on the history of violence. Annual Review of
and Gynaecology. London, UK: Churchill Anthropology 30:573–596.
Livingstone. Walker PL. 2001b. Is the battered child syndrome a
Walker PL. 1969. The linear growth of long bones modern phenomenon? In: Schultz M, editor.
in Late Woodland Indian children. Proceedings Proceedings of the Xth European Meeting of the
of the Indiana Academy of Science 78:83–87. Paleopathology Association Göttingen,
Walker PL. 1978. A quantitative analysis of dental Germany; p. 33–36.
attrition rates in the Santa Barbara Channel Walker PL. 2004. Caring for the dead: finding a
area. American Journal of Physical common ground in disputes over museum
Anthropology 48:101–106. collections of human remains. In: Grupe G,
Walker PL. 1985. Anemia among prehistoric Peter J, editors. Documenta Archaeobiologiae
Indians of the American Southwest. In: Merbs 2: Conservation Policy in Current Research.
CF, Miller RJ, editors. Health and Disease in the Yearbook of the State Collection of
Prehistoric Southwest. Arizona State Anthropology and Palaeoanatomy. Rahden,
University Anthropological Research Papers, Germany: M. Leidorf; p. 13–27.
No. 34; p. 139–164. Walker PL. 2005. Greater sciatic notch
Walker PL. 1986. Porotic hyperostosis in a marine- morphology: sex, age, and population
dependent California Indian population. differences. American Journal of Physical
American Journal of Physical Anthropology Anthropology 127:385–391.
69:345–354. Walker PL. 2006. Skeletal biology: California. In:
Walker PL. 1989. Cranial injuries as evidence of Ubelaker DH, editor. Handbook of North
violence in prehistoric southern California. American Indians: 3. Environment, Origins,
American Journal of Physical Anthropology and Population. Washington, DC: Smithsonian
80:313–323. Institution Press; p. 548–556.
Walker PL. 1990. Tool marks on skeletal remains Walker PL. 2008. Bioarchaeological ethics: a
from Saunaktuk (NgTN-1). In: Arnold CD, historical perspective on the value of human
editor. Archaeological Investigations at remains. In: Katzenberg AM, Saunders SR,
Saunaktuk. Yellowknife, NWT, Canada: Prince editors. Biological Anthropology of the Human
of Wales Northern Heritage Centre, Department Skeleton. New York, NY: John Wiley & Sons;
of Culture and Communications, Government p. 3–40.
of the Northwest Territories; p. 114–123. Walker PL, Bathurst RR, Richman R, Gjerdrum T,
Walker PL. 1991–1992. An overview of California Andrushko VA. 2009. The causes of porotic
Indian history before the arrival of Europeans. hyperostosis and cribra orbitalia: a reappraisal
Journal of Human Ecology 2:359–370. of the iron-deficiency–anemia hypothesis.
Walker PL. 1995. Problems of preservation and American Journal of Physical Anthropology
sexism in sexing: some lessons from historical 139:109–125.
collections for palaeodemographers. In: Walker PL, Byock J, Eng JT, et al. 2011. The
Saunders SR, Herring A, editors. Grave axed man of Mosfell: skeletal evidence of a
Reflections: Portraying the Past through Viking Age homicide, the Icelandic sagas,
Cemetery Studies. Toronto, ON: Canadian and feud. In: Stodder ALW, Palkovich AM,
Scholars’ Press; p. 31–47. editors. The Bioarchaeology of Individuals.
Walker PL. 1997. Wife beating, boxing, and Gainesville, FL: University Press of Florida;
broken noses: skeletal evidence for the cultural p. 26–43.
patterning of interpersonal violence. In: Martin Walker PL, Cook DC, Lambert PM. 1997. Skeletal
D, Frayer D, editors. Troubled Times: Violence evidence for child abuse: a physical
and Warfare in the Past. Langhorne, PA: anthropological perspective. Journal of
Gordon & Breach; p. 145–179. Forensic Sciences 42:196–207.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
584 References

Walker PL, Dean G, Shapiro P. 1991. Estimating University of Oregon Anthropological Papers,
age from tooth wear in archaeological No. 54; p. 355–386.
populations. In: Kelley MA, Larsen CS, editors. Walker PL, Thornton R. 2002. Health, nutrition,
Advances in Dental Anthropology. New York, and demographic change in Native California.
NY: Wiley-Liss; p. 169–178. In: Steckel RH, Rose JC, editors. The Backbone
Walker PL, DeNiro MJ. 1986. Stable nitrogen and of History: Health and Nutrition in the Western
carbon isotope ratios in bone collagen as Hemisphere. New York, NY: Cambridge
indices of prehistoric dietary dependence on University Press; p. 506–523.
marine and terrestrial resources in southern Walker RA, Lovejoy CO, Meindl RS. 1994.
California. American Journal of Physical Histomorphological and geometric properties
Anthropology 71:51–61. of human femoral cortex in individuals over
Walker PL, Erlandson JM. 1986. Dental evidence 50: implications for histomorphological
for prehistoric dietary change on the northern determination of age-at-death. American
Channel Islands, California. American Journal of Human Biology 6:659–667.
Antiquity 51:375–383. Wallace DC, Torroni A. 1992. American Indian
Walker PL, Hewlett BS. 1990. Dental health, diet prehistory as written in the mitochondrial
and social status among Central African DNA: a review. Human Biology 64:403–416.
foragers and farmers. American Anthropologist Wanner IS, Sierra Sosa T, Alt KW, Tiesler Blos V.
92:382–398. 2007. Lifestyle, occupation, and whole bone
Walker PL, Hollimon SE. 1989. Changes in morphology of the pre-Hispanic Maya coastal
osteoarthritis associated with the development population from Xcambo, Yucatan, Mexico.
of a maritime economy among southern International Journal of Osteoarchaeology
California Indians. International Journal of 17:253–268.
Anthropology 4:171–183. Wapler U, Crubezy E, Schultz M. 2004. Is cribra
Walker PL, Johnson J, Lambert P. 1988. Age and orbitalia synonymous with anemia? Analysis
sex biases in the preservation of human and interpretation of cranial pathology in
remains. American Journal of Physical Sudan. American Journal of Physical
Anthropology 76:183–188. Anthropology 123:333–339.
Walker PL, Lambert P. 1989. Skeletal evidence for Ward CV, Latimer B. 2005. Human evolution and
stress during a period of cultural change in the development of spondylolysis. Spine
prehistoric California. In: Capasso L, editor. 30:1808–1814.
Advances in Paleopathology. Journal of Ward T. 1965. Correlations of Mississippian sites
Paleopathology, Monographic Publications, and soil types. Southeastern Archaeological
No.1; p. 207–212. Conference Bulletin 3:42–48.
Walker PL, Lambert P, DeNiro MJ. 1989. The Warinner C, Matias Rodrigues JF, Vyax R, et al. 2014.
effects of European contact on the health of Pathogens and host immunity in the ancient
Alta California Indians. In: Thomas DH, editor. human oral cavity. Nature Genetics 46:336–344.
Columbian Consequences, Volume 2: Warinner C, Tuross M. 2010. Tissue isotopic
Archaeological and Historical Perspectives on enrichment associated with growth depression
the Spanish Borderlands East. Washington, in a pig: implications for archaeology and
DC: Smithsonian Institution Press; ecology. American Journal of Physical
p. 349–364. Anthropology 141:486–493.
Walker PL, Sugiyama L, Chacon R. 1998. Diet, Washburn SL. 1947a. The effect of the temporal
dental health, and cultural change among muscle on the form of the mandible. Journal of
recently contacted South American Indian Dental Research 21:174.
hunter-horticulturalists. In: Lukacs JR, editor. Washburn SL. 1947b. The relation of the temporal
Human Dental Development, Morphology, and muscle to the form of the skull. Anatomical
Pathology: A Tribute to Albert A. Dahlberg. Record 99:239–248.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 585

Washburn SL. 1953. The new physical on the North Carolina coast. In: Powell ML,
anthropology. Yearbook of Physical Cook DC, editors. The Myth of Syphilis: The
Anthropology 7:124–130. Natural History of Treponematosis in North
Watkins R. 2012. Variation in health and America. Gainesville, FL: University Press of
socioeconomic status within the W. Montague Florida; p. 77–91.
Cobb skeletal collection: degenerative joint Webb S. 1989. Prehistoric Stress in Australian
disease, trauma and cause of death. Aborigines: A Palaeopathological Study of a
International Journal of Osteoarchaeology Hunter-Gatherer Population. British
22:22–44. Archaeological Reports, International Series,
Watson JT. 2010. The introduction of agriculture No. 490.
and the foundation of biological variation in Webb S. 1995. Palaeopathology of Aboriginal
the southern Southwest. In: Auerbach BM, Australians: Health and Disease across a
editor. Human Variation in the Americas: the Hunter-Gatherer Continent. Cambridge, UK:
Integration of Archaeology and Biological Cambridge University Press.
Anthropology. Center for Archaeological Webb SG. 1988. Two possible cases of
Investigations, Occasional Paper, No. 33; trephination from Australia. American Journal
p. 135–171. of Physical Anthropology 75:541–548.
Watson JT, Muñoz Ovalle I, Arriaza BT. 2010. Webb SG, Edwards PC. 2002. The Natufian human
Formative adaptations, diet, and oral health in skeletal remains from Wadi Hammeh 27
the Azapa Valley of northwest Chile. Latin (Jordan). Paleorient 28:103–124.
American Antiquity 21:423–439. Webb WS. 1974. Indian Knoll. (Reprinted Edition).
Watt DG, Williams CH. 1951. The effects of the Knoxville, TN: University of Tennessee Press.
physical consistency of food on the growth and Webb WS, Snow CE. 1945. The Adena People.
development of the mandible and maxilla of University of Kentucky Reports in
the rat. American Journal of Orthodontics Anthropology and Archaeology, No. 6.
37:895–928. Weber AW, Katzenberg MA, Schurr TG, editors.
Watts C, White CD, Longstaffe FJ. 2011. 2010. Prehistoric Hunter-Gatherers of the
Childhood diet and its implications for Ontario Baikal Region, Siberia. Philadelphia, PA:
Western Basin subsistence economy at the University of Pennsylvania Press.
Krieger site. American Antiquity 76:446–473. Weber DJ. 1992. The Spanish Frontier in North
Watts R. 2011. Non-specific indicators of stress America. New Haven, CT: Yale University
and their association with age at death in Press.
Medieval York: using stature and vertebral Weber J, Czarnetzki A. 2001. Neurotraumatological
neural canal size to examine the effects of aspects of head injuries resulting from
stress occurring during different periods of sharp and blunt force in the early Neolithic
development. International Journal of period of southwestern Germany. American
Osteoarchaeology 21:568–576. Journal of Physical Anthropology
Waugh LM. 1937. Dental observations among the 114:352–356.
Eskimos. Journal of Dental Research Wedel A, Carlsson GE, Sagne S. 1978.
16:355–356. Temporomandibular joint morphology in
Weaver DS. 1985. Subsistence and settlement Medieval skull material. Swedish Dental
patterns at Casas Grandes, Chihuahua, Mexico. Journal 2:177–187.
In: Merbs CF, Miller RJ, editors. Health and Wegner G. 1874. Ueber das normale und
Disease in the Prehistoric Southwest. Arizona pathologische Wachsthum der Röhrenknocken.
State University Anthropological Research Eine kritische Untersuchung auf
Papers, No. 34; p. 119–127. experimenteller und casiusticher Grundlage.
Weaver DS, Sandford MK, Bogdan G, Kissling GE, Archives für Pathologie und Anatomie
Powell ML. 2005. Prehistoric treponematosis 61:44–76.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
586 References

Weidenreich F. 1943. The skull of Sinanthropus Washington, DC: Smithsonian Institution


pekinensis: a comparative study of a primitive Press; p. 197–225.
hominid skull. Paleontological Sinica 10(whole Wells C. 1975. Prehistoric and historical changes
series):127. in nutritional diseases and associated
Weidenreich F. 1945. The brachycephalization of conditions. Progress in Food and Nutrition
recent mankind. Southwestern Journal of Science 1:729–779.
Anthropology 1:1–54. Wells C. 1980. The human bones. In: Wade-
Weise S, Boldsen JL, Gampe J, Milner GJ. 2009. Martins P, editor. Excavations at North
Calibrated expert inference and construction of Elmham Park 1967–1972. East Anglican
unbiased paleodemographic mortality profiles. Archaeological Report, No. 9; p. 247–374.
American Journal of Physical Anthropology Wells C. 1982. The human burials. In: McWhirr A,
Supplement 48:269. Viner L, Wells C, editors. Cirencester
Weisensee KE. 2013. Assessing the relationship Excavations II: Romano-British Cemeteries at
between fluctuating asymmetry and cause of Cirencester. Cirencester, UK: Cirencester
death in skeletal remains: a test of the Excavation Committee, Corinium Museum;
developmental origins of health and disease p. 135–202.
hypothesis. American Journal of Human Wendell S, Wang X, Brown M, et al. 2010. Tooth
Biology 25:411–417. genes associated with dental caries. Journal of
Weisensee KE, Jantz RL. 2011. Secular changes in Dental Research 89:1198–1202.
craniofacial morphology of the Portuguese Wenham SJ. 1989. Anatomical interpretations of
using geometric morphometrics. American Anglo-Saxon weapon injuries. In: Chadwick
Journal of Physical Anthropology Hawkes S, editor. Weapons and Warfare in
145:548–559. Anglo-Saxon England. Oxford Committee for
Weiss E. 2003. Effects of rowing on humeral Archaeology Monograph, 21; p. 123–139.
strength. American Journal of Physical Wescott DJ. 2006. Effect of mobility on femur
Anthropology 121:293–302. midshaft external shape and robusticity.
Weiss E. 2007. Muscle markers revisited: activity American Journal of Physical Anthropology
pattern reconstruction with controls in a 130:201–213.
central California Amerind population. Wescott DJ, Cunningham DL. 2006. Temporal
American Journal of Physical Anthropology changes in Arikara humeral and femoral cross-
133:931–940. sectional geometry associated with
Weiss E. 2009. Spondylolysis in a pre-contact horticultural intensification. Journal of
San Francisco Bay population: behavioural Archaeological Science 33:1022–1036.
and anatomical sex differences. International Wesolowsky AB. 1989. Osteological analysis. In:
Journal of Osteoarchaeology 19:375–385. Elia RJ, Wesolowsky AB, editors.
Weiss KM. 1972. On the systematic bias in skeletal Archaeological Excavations at the Uxbridge
sexing. American Journal of Physical Almshouse Burial Ground in Uxbridge,
Anthropology 37:239–249. Massachusetts. Boston, MA: Office of Public
Weiss KM. 1973. Demographic Models for Archaeology, Boston University, Report of
Anthropology. Memoirs of the Society for Investigations, No. 76; p. 173–302.
American Archaeology, No. 27. Wessen G, Ruddy FH, Gustafson CE, Irwin H.
Weiss RL, Trihart AH. 1960. Between-meal eating 1977. Characterization of archaeological bone
habits and dental caries experience in pre- by neutron activation analysis. Archaeometry
school children. American Journal of Public 19:200–205.
Health 50:1097–1104. Wessen G, Ruddy FH, Gustafson CE, Irwin H.
Welch PD. 1990. Mississippian emergence 1978. Trace element analysis in the
in west-central Alabama. In: Smith BD, characterization of archaeological bone. In:
editor. The Mississippian Emergence. Carter GF, editor. Archaeological Chemistry – II.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 587

Pullman, IL: American Chemical Society; White CD, Armelagos GJ. 1997. Osteopenia and
p. 99–108. stable isotope ratios of bone collagen of
Weston DA. 2008. Investigating the specificity of Nubian female mummies. American Journal of
periosteal reactions in pathology museum Physical Anthropology 103:185–200.
collections. American Journal of Physical White CD, Healy PF, Schwarcz HP. 1993. Intensive
Anthropology 137:48–59. agriculture, social status, and Maya diet at
Weston DA. 2012. Nonspecific infection in Pacbitun, Belize. Journal of Anthropological
paleopathology: interpreting periosteal Research 49:347–375.
reactions. In: Grauer A, editor. A Companion to White CD, Longstaffe FJ, Law KR. 2004. Exploring
Paleopathology. Chichester, UK: Wiley- the effects of environment, physiology and diet
Blackwell; p. 492–512. on oxygen isotope ratios in ancient Nubian
Whalen RT, Carter DR, Steele CR. 1988. Influence bone and teeth. Journal of Archaeological
of physical activity on the regulation of bone Science 31:233–250.
density. Journal of Biomechanics 21:825–837. White CD, Longstaffe FJ, Schwarcz HP. 2006.
Wheat MM. 1967. Survival Arts of the Primitive Social directions in the isotopic anthropology
Paiutes. Reno, NV: University of Nevada Press. of maize in the Maya region. In: Staller J, Tykot
Wheeler REM. 1968. The Indus Civilization. R, Benz B, editors. Histories of Maize.
Cambridge, UK: Cambridge University Press. Amsterdam, the Netherlands: Elsevier;
Wheeler SM, Williams L, Beauchesne P, Dupras TL. p. 143–159.
2013. Shattered lives and broken childhoods: White CD, Pendergast DM, Longstaffe FJ. 2001.
evidence of physical child abuse in ancient Social complexity and food systems at Altun
Egypt. International Journal of Paleopathology Ha, Belize: the isotopic evidence. Latin
3:71–82. American Antiquity 12:371–394.
White CD. 1988. The ancient Maya from Lamanai, White CD, Price TD, Longstaffe FJ. 2007.
Belize: diet and health over 2,000 years. Residential histories of the human sacrifices at
Canadian Review of Physical Anthropology the Moon Pyramid, Teotihuacan: evidence
6:1–20. from oxygen and strontium isotopes. Ancient
White CD. 1993. Isotopic determination of Mesoamerica 18:159–172.
seasonality in diet and death from Nubian White CD, Schwarcz HP. 1989. Ancient Maya diet
mummy hair. Journal of Archaeological at Lamanai, Belize: as inferred from isotopic
Science 20:657–666. and chemical analysis of human bone. Journal
White CD. 1994. Dietary dental pathology and of Archaeological Science 16:451–474.
cultural change in the Maya. In: Herring A, White CD, Schwarcz HP. 1994. Temporal trends in
Chan L, editors. Strength in Diversity: A Reader stable isotopes for Nubian mummy tissues.
in Physical Anthropology. Toronto, ON: American Journal of Physical Anthropology
Canadian Scholars’ Press; p. 279–302. 93:165–187.
White CD. 1997. Ancient diet at Lamanai and White CD, Spence MW, Longstaffe FJ, Law KR.
Pacbitun: implications for the ecological model 2004. Demography and ethnic continuity in the
of collapse. In: Whittington SL, Reed DM, Tlailotlacan enclave of Teotihuacan: the
editors. Bones of the Maya: Studies of Ancient evidence from stable oxygen isotopes.
Skeletons. Washington, DC: Smithsonian Journal of Anthropological Archaeology
Institution Press; p. 171–180. 23:385–403.
White CD, editor. 1999. Reconstructing Ancient White CD, Wright LE, Pendergast DM. 1994.
Maya Diet. Salt Lake City, UT: University of Biological disruption in the early Colonial
Utah Press. period at Lamanai. In: Larsen CS, Milner GR,
White CD. 2005. Gendered food behaviour among editors. In the Wake of Contact: Biological
the Maya: time, place, status and ritual. Journal Responses to Conquest. New York, NY:
of Social Archaeology 5:356–382. Wiley-Liss; p. 135–145.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
588 References

White TD. 1978. Early hominid enamel Whittington SL. 1999. Caries and antemortem
hypoplasia. American Journal of Physical tooth loss at Copán: implications for
Anthropology 49:79–84. commoner diet. In: White CD, editor.
White TD. 1986. Cut marks on the Bodo cranium: a Reconstructing Ancient Maya Diet. Salt Lake
case of prehistoric defleshing. American City, UT: University of Utah Press; p. 151–167.
Journal of Physical Anthropology 69:503–509. Whittington SL, Reed DM. 1997. Commoner diet at
White TD. 1992. Prehistoric Cannibalism at Copán: insights from stable isotopes and
Mancos 5MTUMR-2346. Princeton, NJ: porotic hyperostosis. In: Whittington SL, Reed
Princeton University Press. DM, editors. Bones of the Maya: Studies of
White TD, Black MT, Folkens PA. 2012. Human Ancient Skeletons. Washington, DC:
Osteology, Third Edition San Diego, CA: Smithsonian Institution Press; p. 157–170.
Academic Press. Whittle A. 1996. Europe in the Neolithic.
White TD, Degusta D, Richards GD, Baker SG. Cambridge, UK: Cambridge University Press.
1997. Prehistoric dentistry in the American Widmer L, Perzigian AJ. 1981. The ecology and
Southwest: a drilled canine from Sky Aerie, etiology of skeletal lesions in late prehistoric
Colorado. American Journal of Physical populations of eastern North America. In:
Anthropology 103:409–414. Buikstra JE, editor. Prehistoric Tuberculosis in
Whitehorne J. 1991. Fort Erie and U.S. operations the Americas. Center for American Archeology,
on the Niagara frontier, 1814. In: Pfeiffer S, Scientific Papers, No. 5; p. 99–113.
Williamson RF, editors. Snake Hill: An Wiechmann I, Grupe G. 2005. Detection of
Investigation of a Military Cemetery from the Yersinia pestis DNA in two early Medieval
War of 1812. Toronto, ON: Dundurn Press; skeletal finds from Aschheim (Upper Bavaria,
p. 25–60. 6th century A.D.). American Journal of
Whitley AT, Kendrick GS, Mathews JL. 1966. The Physical Anthropology 126:48–55.
effects of function on osseous and muscle Wilbur AK, Bouwman AS, Stone AC, et al. 2009.
tissues in the craniofacial area of the rat. Angle Deficiencies and challenges in the study of
Orthodontist 36:13–17. ancient tuberculosis DNA. Journal of
Whitney EN, Rolfes SR. 2011. Understanding Archaeological Science 36:1990–1997.
Nutrition, Twelfth Edition. Belmont, CA: Wilbur AK, Farnbach AW, Knudson KJ, Buikstra
Cengage Learning. JE. 2008. Diet, tuberculosis, and the
Whittaker DK. 1993. Oral health. In: Molleson T, paleopathological record. Current
Cox M, editors. The Spitalfields Project: Anthropology 49:963–991.
Volume 2 – The Anthropology. The Middling Wilbur AK, Stone AC. 2012. Using ancient DNA
Sort. Council for British Archaeology Report, techniques to study human disease. In:
No. 6; p. 49–65. Buikstra JE, Roberts C, editors. The Global
Whittaker DK, Molleson T, Daniel AT, et al. 1985. History of Paleopathology: Pioneers and
Quantitative assessment of tooth wear, Prospects. New York, NY: Oxford University
alveolar-crest height and continuing eruption Press; p. 703–717.
in a Romano-British population. Archives of Wilczak C, Watkins R, Null CC, Blakey ML. 2009.
Oral Biology 30:493–501. Skeletal indicators of work: musculoskeletal,
Whittaker DK, Richards D. 1978. Scanning electron arthritic, and traumatic effects. In: Blakey ML,
microscopy of the neonatal line in human Rankin-Hill LM, editors. The Skeletal Biology of
enamel. Archives of Oral Biology 23:45–50. the New York African Burial Ground, Part I.
Whittington SL. 1989. Characteristics of Washington, DC: Howard University Press;
Demography and Disease in Low-Status Maya p. 198–226.
from Classic Period Copán, Honduras. PhD Wilkinson RG, Norelli RJ. 1981. A biocultural
Dissertation, Pennsylvania State University, analysis of social organization at Monte Albán.
University Park, State College, PA. American Antiquity 46:743–758.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 589

Wilkinson RG, Van Wagenen KM. 1993. Violence general health: a consensus view. Current
against women: prehistoric skeletal evidence Medical Research and Opinion 24:1635–1643.
from Michigan. Midcontinental Journal of Williamson MA. 2000. A comparison of
Archaeology 18:190–216. degenerative joint disease between upland and
Willerslev E, Cooper A. 2005. Ancient DNA. coastal prehistoric agriculturalists from
Proceedings of the Royal Society B 272:3–16. Georgia. In: Lambert PM, editor.
Willey P. 1990. Prehistoric Warfare on the Great Bioarchaeological Studies of Life in the Age of
Plains: Skeletal Analysis of the Crow Creek Agriculture: A View from the Southeast.
Massacre Victims. New York, NY: Garland Tuscaloosa, AL: University of Alabama Press;
Publishing. p. 134–147.
Willey P, Emerson TE. 1993. The osteology and Williamson RF, Pfeiffer S. 2003. Bones of the
archaeology of the Crow Creek massacre. In: Ancestors: The Archaeology and
Tiffany JA, editor. Prehistory and Human Osteobiography of the Moatfield Ossuary.
Ecology of the Western Prairies and Northern Canadian Museum of Civilization, Mercury
Plains: Papers in Honor of Robert A. Alex Series Archaeology Paper, No. 163.
(1941–1988). Plains Anthropologist Memoir, Willis A, Oxenham MF. 2013. The Neolithic
No. 27; p. 227–269. demographic transition and oral health: the
Willey P, Scott DD. 1996. “The bullets buzzed like Southeast Asian experience. American
bees”: gunshot wounds in skeletons from the Journal of Physical Anthropology
Battle of the Little Bighorn. International 152:197–208.
Journal of Osteoarchaeology 6:15–27. Wilson DE. 2005. Treponematosis in the East
Williams JA. 1994. Disease profiles of Archaic and Texas Gulf coastal plain. In: Powell ML, Cook
Woodland populations in the northern Plains. DC, editors. The Myth of Syphilis: the Natural
In: Owsley DW, Jantz RL, editors. Skeletal History of Treponematosis in North America.
Biology in the Great Plains: Migration, Gainesville, FL: University Press of Florida;
Warfare, Health, and Subsistence. Washington, p. 162–176.
DC: Smithsonian Institution Press; p. 91–108. Wilson DJ. 1987. Reconstructing patterns of early
Williams JA, Snortland-Coles JS. 1986. Pre- warfare in the lower Santa Valley: new data on
contact tuberculosis in a Plains Woodland the role of conflict in the origins of complex
mortuary. Plains Anthropologist 31:249–252. north-coast society. In: Haas J, Pozorski S,
Williams JS, White CD, Longstaffe FJ. 2005. Pozorski T, editors. The Origins and
Trophic level and macronutrient shift effects Development of the Andean State. Cambridge,
associated with the weaning process in the UK: Cambridge University Press; p. 56–69.
Postclassic Maya. American Journal of Wiltse LL. 1957. Etiology of spondylolisthesis.
Physical Anthropology 128:781–790. Clinical Orthopaedics 10:48–58.
Williams JS, White CD, Longstaffe FJ. 2009. Maya Wiltse LL. 1962. The etiology of spondylolisthesis.
marine subsistence: isotopic evidence for trade Journal of Bone and Joint Surgery
and status from Marco Gonzalez and San 44-A:539–560.
Pedro, Belize. Latin American Antiquity Wiltse LL, Widell EH, Jackson DW. 1975. Fatigue
20:37–56. fracture: the basic lesion in isthmic
Williams LL. 2009. “The dam is becoming spondylolisthesis. Journal of Bone and Joint
dangerous and may possibly go”: The Surgery 57-A:17–22.
paleodemography of the Johnstown Flood of Winchell F, Rose JC, Moir RW. 1995. Health and
1889. In: Jones EC, Murphy AD, editors. The hard times: a case study from the middle to late
Political Economy of Hazards and Disasters. nineteenth century in eastern Texas. In: Grauer
Lanham, MD: Altamira Press; p. 31–57. AL, editor. Bodies of Evidence: Reconstructing
Williams RC, Barnett AH, Claffey N, et al. 2008. History through Skeletal Analysis. New York,
The potential impact of periodontal disease on NY: Wiley-Liss; p. 161–172.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
590 References

Winkler LA, Larsen CS, Thompson V, Sciulli PW, Wood JW, Milner GR, Harpending HC, Weiss KM.
Hutchinson DL. 2012. The social structuring of 1992. The osteological paradox: problems in
stress in contact-era Spanish Florida: a inferring prehistoric health from skeletal
bioarchaeological case study from Santa samples. Current Anthropology 33:343–358.
Catalina de Guale, St. Catherines Island, Wood JW, Milner GR. 1994. Reply. Current
Georgia. American Journal of Physical Anthropology 35:631–637.
Anthropology Supplement 54:305. Woodburn J. 1968. An introduction to Hadza
Winterhalder B, Kennett DJ. 2009. Four neglected ecology. In: Lee RB, DeVore I, editors. Man the
concepts with a role to play in explaining the Hunter. Chicago, IL: Aldine; p. 49–55.
origins of agriculture. Current Anthropology Woodward M, Walker ARP. 1994. Sugar
50:645–648. consumption and dental caries: evidence from
Wittmers LE Jr., Aufderheide AC, Pounds JG, 90 countries. British Dental Journal
Jones KW, Angel JL. 2008. Problems in 176:297–302.
determination of skeletal lead burden in World Health Organization. 2007. Global
archaeological samples: an example from the Tuberculosis Control. World Health
First African Baptist Church population. Organization, Geneva, Switzerland.
American Journal of Physical Anthropology Worne H, Cobb CR, Vidoli G, Steadman DW. 2012.
136:379–386. The space of war: connecting geophysical
Wittmers LE Jr., Aufderheide AC, Wallgren J, Rapp landscapes with skeletal evidence of warfare-
G, Alich A. 1988. Lead in bone IV. Distribution related trauma. In: Martin DL, Harrod RP, Pérez
of lead in the human skeleton. Archives of VP, editors. The Bioarchaeology of Violence.
Environmental Health 43:381–391. Gainesville, FL: University Press of Florida;
Wittwer-Backofen U, Buckberry J, Czarnetzki A, p. 141–159.
et al. 2008. Basics in paleodemography: a Worth JE. 1995. The Struggle for the Georgia
comparison of age indicators applied to the Coast: An Eighteenth-Century Spanish
Early Medieval skeletal sample of Lauchheim. Retrospective on Guale and Mocama.
American Journal of Physical Anthropology Anthropological Papers of the American
137:384–396. Museum of Natural History, No. 75.
Wolff J. 1892. The Law of Bone Remodelling. Worth JE. 2001. The ethnohistorical context of
(Translation by P. Maquet & R. Furlong). Berlin, bioarchaeology in Spanish Florida. In: Larsen
Germany: Springer-Verlag. CS, editor. Bioarchaeology of Spanish Florida:
Wolpoff MH. 1971. Interstitial wear. American The Impact of Colonialism. Gainesville, FL:
Journal of Physical Anthropology 34:205–228. University Press of Florida; p. 1–21.
Woo EJ, Sciulli PW. 2013. Degenerative joint Worth JE. 2007. The Struggle for the Georgia Coast.
disease and social status in the terminal Late Tuscaloosa, AL: University of Alabama Press.
Archaic period (1000–500 B.C.) of Ohio. Wright JT. 2010. Defining the contribution of
International Journal of Osteoarchaeology genetics in the etiology of dental caries.
23:529–544. Journal of Dental Research 89:1173–1174.
Woo SLY, Kuei SC, Amiel D, et al. 1981. The effect Wright LE. 1990. Stresses of conquest: a study of
of prolonged physical training on the Wilson bands and enamel hypoplasias in the
properties of long bone: a study of Wolff’s Maya of Lamanai, Belize. American Journal of
Law. Journal of Bone and Joint Surgery Human Biology 2:25–35.
63-A:780–787. Wright LE. 1997. Intertooth patterns of hypoplasia
Wood BF. 1971. Malocclusion in the modern expression: implications for childhood health
Alaskan Eskimo. American Journal of in the Classic Maya collapse. American Journal
Orthodontics 60:344–354. of Physical Anthropology 102:233–247.
Wood CS. 1979. Human Sickness and Health: Wright LE. 2006. Diet, Health, and Status among
A Biocultural View. Palo Alto: Mayfield, CA. the Pasion Maya: A Reappraisal of the

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
References 591

Collapse. Nashville, TN: Vanderbilt University prehistoric Japan. Journal of the


Press. Anthropological Society of Nippon
Wright LE, Schwarcz HP. 1996. Infrared and 90(S):77–90.
isotopic evidence for diagenesis of bone apatite Yamamoto M. 1992. Secular trends of enamel
at Dos Pilas, Guatemala: paleodietary hypoplasia in Japanese from the prehistoric to
implications. Journal of Archaeological Science modern period. In: Goodman AH, Capasso LL,
23:933–944. editors. Recent Contributions to the Study of
Wright LE, Schwarcz HP. 1998. Stable carbon and Enamel Developmental Defects. Journal of
oxygen isotopes in human tooth enamel: Paleopathology, Monographic Publications,
identifying breastfeeding and weaning in No. 2; p. 231–238.
prehistory. American Journal of Physical Yang X, Wan Z, Perry L, et al. 2012. Early
Anthropology 106:1–18. millet use in northern China. Proceedings
Wright LE, Valdes JA, Burton JH, Price TD, of the National Academy of Science
Schwarcz HP. 2010. The children of 109:3726–3730.
Kaminaljuyu: isotopic insight into diet and Yano K, Wasnich RD, Vogel JM, Heilbrun LK.
long distance interaction in Mesoamerica. 1984. Bone mineral measurements among
Journal of Anthropological Archaeology middle-aged and elderly Japanese residents in
29:155–178. Hawaii. American Journal of Epidemiology
Wright LE, White CD. 1996. Human biology in the 119:751–764.
Classic Maya collapse: evidence from y’Edynak G. 1989. Yugoslav Mesolithic dental
paleopathology and paleodiet. Journal of reduction. American Journal of Physical
World Prehistory 10:147–198. Anthropology 78:17–36.
Wu J. 1994. How severe was the Great Depression? y’Edynak G, Fleisch S. 1983. Microevolution and
Evidence from the Pittsburgh region. In: biological adaptability in the transition from
Komlos J, editor. Stature, Living Standards, food-collecting to food-producing in the Iron
and Economic Development: Essays in Gates of Yugoslavia. Journal of Human
Anthropometric History. Chicago, IL: Evolution 12:279–296.
University of Chicago Press; p. 129–152. Yoder C. 2010. Diet in Medieval Denmark: a
Wu X, Zhang Z. 1985. Homo sapiens remains from regional and temporal comparison. Journal of
late Palaeolithic and Neolithic China. In: Archaeological Science 37:2224–2236.
Rukang Wu, Olsen JW, editors. Yoneda M, Suzuki R, Shibata Y, et al. 2004.
Palaeoanthropology and Palaeolithic Isotopic evidence of inland-water fishing by
Archaeology in the People’s Republic of China. Jomon population excavated from the Boji site,
Orlando, FL: Academic Press; p. 107–133. Japan. Journal of Archaeological Science
Wu XJ, Liu W, Bae CJ. 2012. Craniofacial 31:97–107.
variation between southern and northern Zaino EC. 1967. Symmetrical osteoporosis, a sign
Neolithic and modern Chinese. International of severe anemia in the prehistoric Pueblo
Journal of Osteoarchaeology 22:98–109. Indians in the Southwest. In: Wade WD, editor.
Wynne-Davies R, Scott JHS. 1979. Inheritance Miscellaneous Papers in Paleopathology: I.
and spondylolythesis: a radiographic family Museum of Northern Arizona, Technical Series,
survey. Journal of Bone and Joint Surgery No. 7; p. 40–47.
61-B:301–305. Zaino EC. 1968. Elemental bone iron in the
Yagi T, Takebe Y, Itoh M. 1989. Secular trends in Anasazi Indians. American Journal of Physical
physique and physical fitness in Japanese Anthropology 29:433–435.
students during the last 20 years. American Zakrzewski SR. 2007. Population continuity or
Journal of Human Biology 1:581–587. population change: formation of the ancient
Yamaguchi B. 1982. A review of the osteological Egyptian state. American Journal of Physical
characteristics of the Jomon population in Anthropology 132:501–509.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
592 References

Zeder MA. 2008. Domestication and early and Clinical Approaches. British
agriculture in the Mediterranean Basin: origins, Archaeological Reports, International Series,
diffusion, and impact. Proceedings of the No. 1054; p. 259–268.
National Academy of Sciences Zimmerman LJ, Emerson T, Willey P, et al. 1981.
105:11597–11604. The Crow Creek Site (39BF11) Massacre:
Zegura S, Karafet T, Zhivotovsky L, Hammer M. A Preliminary Report. Omaha, NE: US Army
2004. High-resolution SNPs and microsatellite Corps of Engineers.
haplotypes point to a single, recent entry of Zink AR, Grabner W, Nerlich AG. 2005. Molecular
Native American Y chromosomes into the identification of human tuberculosis in recent
Americas. Molecular Biology and Evolution and historic bone tissue samples: the role of
21:164–175. molecular techniques for the study of historic
Zhang F, Zhi X, Jingze T, et al. 2010. Prehistorical tuberculosis. American Journal of Physical
East–West admixture of maternal lineages in a Anthropology 126:32–47.
2,500-year-old population in Xinjiang. Ziskin DE. 1926. The incidence of dental caries in
American Journal of Physical Anthropology pregnant women. American Journal of
142:314–320. Obstetrics and Gynecology 12:710–719.
Zhang Y, Jordan JM. 2008. Epidemiology of Zuckerman MK, Armelagos GJ. 2011. The origins
osteoarthritis. Rheumatic Disease Clinics of of biocultural dimensions in bioarchaeology.
North America 34:515–529. In: Agarwal SC, Glencross BA, editors. Social
Zhou L, Corruccini RS. 1998. Enamel hypoplasias Bioarchaeology. Chichester, UK: Wiley-
related to famine stress in living Chinese. Blackwell; p. 15–43.
American Journal of Human Biology Zumwalt A. 2006. The effect of endurance exercise
10:723–733. on the morphology of muscle attachment sites.
Zias J. 1991. Leprosy and tuberculosis in the Journal of Experimental Biology 209:444–454.
Byzantine monasteries of the Judean Desert. Zumwalt AC, Ruff CB, Wilczak CA. 2000. Primate
In: Ortner DJ, Aufderheide AC, editors. Human muscle insertions: What does size tell you?
Paleopathology: Current Syntheses and Future American Journal of Physical Anthropology
Options. Washington, DC: Smithsonian Supplement 30:331.
Institution Press; p. 197–204. Zvelebil M. 1986. Mesolithic prelude and Neolithic
Zias J. 2002. New evidence for the history of revolution. In: Zevlebil M, editor. Hunters in
leprosy in the ancient Near East: an overview. Transition: Mesolithic Societies of Temperate
In: Roberts CA, Lewis ME, Manchester K, Europe and their Transitions to Farming.
editors. The Past and Present of Leprosy: Cambridge, UK: Cambridge University Press;
Archaeological, Historical, Palaeopathological p. 5–15.

Downloaded from https:/www.cambridge.org/core. University of Florida, on 09 May 2017 at 06:55:13, subject to the Cambridge Core terms of use, available at
https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139020398.014
INDEX

Page numbers for figures are in italics.

abrasion (tooth) 277 and fertility/birth rate 411-418 alignment of fractures 169
Abu Hureyra (Syria) 205, 280, 295 decline with agricultural Amelia Island (Santa Catalina de
accidental injuries 115-117, intensification 414-418,417 Guale de Santa Maria,
127-130, 157, 172 juvenility index 413-414, 414, Florida) 70, 235-237,
forearm (may be due to violence) 415, 416 236, 303
122-124, 123, 124, 157 population pyramids 411, 412 ameloblasts 44
humerus 128 infant mortality 330, 405, America see Mesoamerica; North
pre-existing, in soldiers 161-162, 407-408 America; South America
169 life table analysis 406-409 amino acids 16, 317
rib 117-118, 118 mortality profiles 404-406 Anasazi people (USA) 146-148,
treatment 168-171 problems in age estimation 407, 308, 394
vertebral 118-122, 119 409-410, 418-419, 419 anemia 30-31
Admiralty Island (Alaska) 145 stress indicators 54, 65 cranial porotic lesions 30-41, 32,
adolescent growth rates 9-10 underrepresentation of juveniles 34, 317
Africa 407-408, 413 age distribution 34, 39-41, 41
central African tribes 74-75, variability in preservation of elemental analysis of iron
78, 430 skeletal elements 403 352-353
South Africa 264, 287, 323, aggression see violence/violent genetic 31, 33
370-372, 428 injuries Angel, J. Lawrence 404
see also Nubia agriculturalists Anglo-American War (1812) 167
African-Americans accidental injuries 121, 129-130 Anglo-Saxons 12-13, 162
craniofacial morphology 258 bone infections 88-94 animal skins, chewing 84, 188, 287
during enslavement bone remodeling due to ankylosis 181
lead levels 354 mechanical forces 225, 226, apatite 302-304, 338, 355-356
tooth enamel defects 51 228-229 Arabia 343, 344
osteoarthritis 191-192, 196 dentition archaeological context 3-4, 422-424
periosteal reactions 95 caries 69-73, 70 archers 193, 211, 248
age enamel defects 53-54 Arctic peoples
bone mass/loss 57, 58, 63-64, periodontal disease/tooth loss bone remodeling due to
125-126 81-83 mechanical forces 225-227
bone remodeling 241-246, 244, tooth size 26 craniofacial morphology
245 tooth wear 279-282, 284-287, 260-261, 265-268
C and N isotope ratios 330-332 285, 286, 292-297 dentition 266, 287, 290
cranial porosities 34, 39-41, 41 fertility decline 414-418, 417 damaged teeth 297-298, 299
craniofacial morphology 261, nutritional stress 10-11, 16-18, 37 entheseal changes in the arm
268-270 osteoarthritis 180, 192, 194-196 209-210
dentition pastoralists 54, 253, 370-372 Harris lines 43
caries 327 see also cereal diets; foraging-to- osteoarthritis 188, 191, 193, 196
enamel defects 49, 50, 50-51 farming transition; maize osteon structure 63
tooth loss 82 Ainu (Japan) 376 spondylolysis 119
entheseal changes 210 Aka tribe (Africa) 74-75, 78, 430 violent injuries 143-146
Harris lines 44 Alabama (USA) see Roger’s Island; Argentina 263, 266
osteoarthritis 180-181, 196-197 Moundville; Pickwick Basin; Ankara people (USA)
violent injuries 135-136, 139-140, Tombigbee River valley biodistance analysis 385, 398
155 Alaska (USA) 145-146, 209-210, bone remodeling due to
of weaning 323-330, 325, 326, 225-227, 260-261, 380, mechanical forces 237-238
328, 349-351 388-389 nutritional status 11, 11, 59-60,
age-at-death 402-404, 421 Aleutian islands (Alaska) 209-210, 174-175, 426
disaster events 420-421, 421 225-227, 380 violent injuries 136-138
Index

Arizona (USA) 194, 211, 308, maize consumption 309-310, loading forces 215, 216
350-351 312, 312, 314 mobile lifestyle 250-253, 251,
arm see upper limb Belleville (Ontario) 12-13, 118, 192, 252, 254
arrow injuries 134, 142-143 329 osteoarthritis 232
arthritis see osteoarthritis bending rigidity (/) 220 sex differences
articular joints see joints bending stresses 215, 217,218, foraging-to-farming transition
ascorbic acid (vitamin C) 31 219-220 228-229, 232-233, 238,
scurvy 39, 40 Beringa (Peru) 151 240-241, 241
Asia biodistance analysis mobility 223-224, 251, 252, 253
biodistance analysis 373-377, data types and socioeconomic status in the
378, 379 genetic 359-360, 362, Mesolithic 230
dentition 367-368 socioeconomic status 230, 248
caries 71-73 metric 359, 362-363 soldiers 230, 248
tooth loss 82 nonmetric 359, 360, 361-364, and terrain 225, 253
tooth wear 281, 284, 288, 291, 364, 367 birth rate in paleodemography
296 overview 357-362, 358, 423-424 411-418
rice consumption 71, 281 spatial relationships decline with agricultural
violent injuries 159 ethnic boundaries 389-390, 390 intensification 414-418, 417
see also China; Indian interbreeding between juvenility index 413-414, 414,
subcontinent; Japan; Siberia Europeans and native 415, 416
athletes 214, 243, 250 peoples 399-401 population pyramids 411, 412
atlatls 193-194, 210-211 kinship and marriage 390-396, birth weight, low 7-8
attrition (tooth) 277 393, 395, 397 blunt force trauma 143, 168
Australian Aborigines social distinctions 397-399, 400 see also cranial injuries
cranial porosities 36-37 temporal changes: population Boaz, Franz 257
injuries 124, 157-158 movement and cultural bog corpses 160
treatment 169 transition 368 bone density 216
osteoarthritis 203 into America 377-383, 382 bone mass
periostea] reactions 94 within American regions and cortical area 219, 249, 250
repatriation of remains 428 383-386 decreased
tooth wear 287-288 Central Asia 374-375 in adults 57-60, 58, 63-64, 223
Australopithecus spp. 186-187 East and Southeast Asia and fracture risk 125-126
Averbuch (Tennessee) 94, 195-196 375-377, 378, 379 inactivity-related 215
axial loading 219-220 Europe 372-373 in juveniles 13-14
Ayalan (Ecuador) 205 Indus River valley 373-374 bone remodeling see biomechanics
linking living groups to and skeletal morphology
Bab edh-Dhra’ (Jordan) 205, ancestors 388-389 bows and arrows
391-392 Nubia 369-370 injuries inflicted by 134, 142-143
back see vertebrae South Africa 370-372 upper limb changes associated
Bactrian culture 374 Spanish Florida 361, 386-388, with archery 193, 211, 248
Bahamas 337 387, 388 brachycephalic head shape
Ban Chiang (Thailand) 345-346 biomechanics and skeletal morphology 258-259
Bantu farmers 74-75 214- 215,255,430 Brazil 119, 149, 290-291
Barbados 354 age changes 241-246, 244, 245 breastfeeding 38, 323
barium (Ba) 350-351, 352 agriculturalists 225, 226, 228-229 see also weaning
Barker hypothesis bilateral asymmetry 214, Bronze Age 130, 343, 368, 375
see also developmental origins of 227- 231, 231, 248 see also Indus Valley civilization
disease hypothesis cross-sectional geometry brow ridge (supraorbital torus)
Barlowe, Arthur 279 215- 222, 218, 219 264-265
basketmaking 289-290 as applied to specific groups bubonic plague 112
battered-child syndrome 174, 177 222-246 burial practices
beer (chicha) 76, 313 external measurements 247-255 prone position 162-163
beheading 133, 149, 160, 162-163, femoral neck-shaft angle 253-255 and status 95, 197
167, 175 foragers 222-227, 223, 225, 251 treatment of juveniles 409
trophy heads 152 foraging-to-farming transition butchery marks 145-148, 147, 149
bejel (endemic syphilis) 96-97 228- 229, 231-241, 234,
Belize 236, 239, 241, 247-249 C3 plants, 5°C values 302-303, 303
environmental stress 54-55 histomorphometry 60-64, 61, 62, C4 plants, 813C values 302-303, 303
family burials 392 246-247 see also maize; millet
Index

cacti in the diet 303, 307-308 caries 67, 67 teeth


Cahokia (Illinois) 38, 96, 306 and age at weaning 327 caries 57, 327
calcium 58-59, 63, 348 causes and associations 56-57, enamel defects 48, 52-53
California (USA) 68- 69, 429-430 eruption 25, 76
dentition 55, 73, 282 and diet 68, 313-315, 320, size 26-28
dietary shift 19, 73, 282, 336-337 430-431 see also weaning
environmental stress sex differences 73-77 Chile
cranial porosities 36 socioeconomic status 77-78, 313 craniofacial robusticity
dental conditions 55 temporal trends in prevalence (Patagonia) 266
Harris lines 43 69- 73, 70 elite males in the Maitas-
stature reduction 19 and tooth size 274-276 Chiribaya culture 19, 96
osteoarthritis 202-203 treatment 170-171, 171 tuberculosis 104-105, 107-108
periosteal reactions 93-95 caries sicca 97, 99, 101, 101 violent injuries 154
violent injuries 123, 140-143, carnivores, stable isotope analysis China
144 320, 348, 350 dental enamel hypoplasia during
CAM (crassulacean acid metabolism) Carrier Mills (Illinois) 350 modem famine 46, 47
photosynthesis 303 Carthage (Tunisia) 353 mandibular tori 268
Cambodia 159 fatalhoyiik (Turkey) millet cultivation 317
Canada biodistance analysis of kinship stature decline related to
Northwest Territories 144-145, 188 patterns 394-396 agriculture IS, 18
Ontario see Ontario bone rigidity/strength 244, 245 Tarim Basin mummies 375
cancellous (trabecular) bone 63-64 diet 321, 322 tooth size increase 273
cannibalism 145-148, 147, 149, fractures 125-126, 126 chipped teeth 297-300, 298, 299
155-156, 173 growth rates 11 Chiribaya culture (Pem) 96, 336
canoeing 188, 205, 209, 225-227 mobility 253, 254 chopsticks 281
Canzo, Gonzalo Mendez de, paleodemography 405, 408-409 Chucalissa (Tennessee) 130
Governor of Spanish Florida weaning 325, 326 circular caries 56-57
200 cereal diets 425 climate
carbohydrate consumption anemia 34-35, 37, 317 and body/head shape 15, 266
and caries 68, 430-431 caries 68, 70-71 and Harris lines in Arctic
see also cereal diets; maize maize see maize peoples 43
carbon (C), analysis of stable isotope millet 59-60, 316-318, 317 815N values 322-323
ratios rice 71, 281 oxygen isotope ratios 346
and dietary sources of carbon sorghum 318 and rickets 21
302-304, 303 Cerro Cerrillos (Peru) 152 climate change 431-432
maize consumption cervical spine American Pacific area 142
comparison with marine diet injuries 121-122, 127-129 Little Ice Age 346
333-336, 334, 335 osteoarthritis 195-196 and maize production 308, 309
in Euroamericans 315-316 Chalcatzingo (Mexico) 349 and violence 173, 177
in Mesoamerica 308-310, 312, Chavez Pass (Arizona) 194 cold adaptation and craniofacial
312- 314 chewing see dentition, morphology 265-266
in North America 304, extramasticatory uses; collagen 302-304, 338, 355-356
304-308, 305, 309, 313 masticatory behavior Colies’ fractures 122, 124
sex differences 313-315, 314 children Colorado (USA) 146-148
socioeconomic status 306, 312, bone rigidity/strength 242-243, computed tomography (CT) 221-222
313- 315 244, 245 Conchopata (Pem) 152
in South America 310-311, child abuse 174, 177 congenital syphilis 97, 100
311, 313 environmental stress 7-8, 54, 64 context in bioarchaeology 3-4,
marine diet 304, 310 cranial porosities 39-41 422-424
bivariate 613C/815N analysis growth rates 9-14, 11, 64 cooking 260, 267-268, 280, 354
333-338, 334, 335 rickets 21-23, 23 Copan (Honduras) 57, 310,
and the foraging-to-farming teeth 25-28, 48, 52-53 312-314
transition 318-320, 319 infant mortality 330,405, 407-408 cortical area (CA) 219, 249, 250
millet consumption 316-318, 317 lead toxicity 354 cortical bone mass
carbonate, in bone apatite 302-304, paleodemography and environmental stress 13-14,
338, 355-356 juvenility index 413-414, 414, 57-60, 58
cardiovascular disease 85 415, 416, 417 and fracture risk 125-126
Caribbean islands 337, 354 problems with 407-409, 413 and physical inactivity 215
596 Index

Costa Rica 309 humeral bilateral asymmetry notched teeth 290


Cowboy Wash (Colorado) 148 227-231, 231 isotope analysis 302, 356
cranial base height 20-21 Crow Creek (South Dakota) 136-137, oxygen 346-347
cranial injuries 117 174-175 strontium 339-346, 340, 344,
accidental 127-129 345
Australia 157-158 Dakhleh Oasis (Egypt) 92-93 leprosy 111
Cambodia 159 De Poncins, G. 265 malocclusion 270-273
in conjunction with parry decapitation 133, 149, 160, periodontal disease and lost teeth
fractures 122-123 162-163, 167, 175 78-81, 79, 82
Easter Island 155 trophy heads 152 in pregnancy 76
Europe 159, 161 deciduous teeth sex differences 83-85, 84
North America 133, 133, 135-137, caries 57, 327 socioeconomic status 84-85
140-143, 141, 166, 168 enamel defects 48, 52-53 and systemic health 85-86
skull position 140-142, 157-158, eruption 25 temporal trends in prevalence
174 tooth size 26-28 81-83
South America 151, 151-152 degenerative conditions 6, 29, 85 shovel-shaped incisors 266
treatment 169-170, 170 see also caries; osteoarthritis; size of teeth 26-28, 49, 271-276,
cranial porosities spondylolysis 274
age profile 34, 39-41, 41 demography see paleodemography syphilis (congenital) 97, 100
and anemia 30-41, 32, 34, 317 Denmark treatment of diseased teeth
and scurvy 39, 40 diet 318, 319 170-171, 171
cranial shape variation in biodistance leprosy 109-110 wear on teeth 276-278
analysis 363,364 mercury 354-355 buccal surface 297
Americas 382-383, 385-386, 391 Trelleborg soldiers 344-345, 345 and caries 68-69
Asia 375, 377, 379 violent injuries 159-162, 169 and diet 277-288, 281, 292,
Easter Island 399-400 dentition 292-297
Europe 373 asymmetry due to extramasticatory uses
and kinship 390-391, 394 directional 29 287-292, 289, 291
Nubia 369-370 fluctuating 28-29, 30 interproximal 282-283
South Africa 371 biodistance analysis 359, 360, lingual surface 290-292, 291
cranial vault lesions in 361, 363-364, 367 lingual tilting of molar 278
treponematosis 97, 99, 101, Americas 380,384,388-389,399 occlusal surface macrowear
101 East Asia 375-376, 378 patterns 283-287, 285,
craniofacial morphology 256 Indus River valley 374 286
and age 261, 268-270 kinship and marriage 391-392, occlusal surface macrowear
Arctic peoples 260-261, 265-268 393, 394-396 severity 277-282, 281
factors determining form 256-257 Nubia 369-370 occlusal surface microwear
malocclusion 270-273 South Africa 371 292, 292-297
masticatory behavior 258, caries see caries sex differences 287-290
260-268, 261, 262 damaged teeth 297-300, 298, 299 socioeconomic status 288,
oral hard tissue tori 266-268, 267 development rate 25, 76 296-297
and race 257-259 enamel defects 44-57, 45 desert habitats
supraorbital torus 264-265 age at which stress occurred 49, Australia, health indicators
temporal trends 257-264, 258, 50, 50-51 36-37, 94
261, 267, 272-273 and caries 56-57 Ba/Sr ratios 351
tooth size changes 271-276, 274 duration/severity of stress 51-52 815N values 322-323
crassulacean acid metabolism (CAM) hypocalcification 46, 56-57 Saharan pastoralism 54
photosynthesis 303 hypoplasia 44-48, 45, 47, developmental origins of disease
cribra orbitalia 33-34, 34, 39-41 52- 57, 56 hypothesis 7-8, 54, 64, 431
Cross family (American pioneers) Wilson bands 48, 48-49, 51, diaphyseal bowing 22-23, 23
315, 391 53- 55 diaphyseal flattening 23-24, 249
cross-sectional geometry of bone extramasticatory uses 288-292 diarrhea 34, 37-38
215-222, 218, 219 chewing hides 84, 188, 287 Dickson Mounds (Illinois)
and age 241-246, 244, 245 chipped teeth 298 environmental stress 24, 44, 54
in foragers 222-227, 223, 225 grooved teeth 288-290, 289 fractures 130
foraging-to-farming transition Inuit 84, 188, 265, 287, 290 infections 89-90
228-229, 231-241, 234, lingual surface wear 290-292, multi-element analysis 352
236, 239, 241 291 osteoarthritis 196, 202
Index 597
V.

diet vitamin B3 16 Wilson bands 48, 48-49, 51,


barium 350-351, 352 vitamin B12 32, 38 53-55
and caries 68, 320, 430-431 vitamin C 31, 39 endemic syphilis (bejel) 96-97
comparison with isotope ratios vitamin D 21-23, 23 endogamy 391-392
313-315 weaning see weaning England
sex differences 74-76 zinc 352-353 cranial porotic lesions 35
shift to agriculture 69-73, 70 directional asymmetry in dentition 29 craniofacial morphology 258
socioeconomic status 77-78, dismemberment 133, 143, 145 growth rates 10, 13, 14, 24-25
313 DNA analysis 362, 367-368 humeral bilateral asymmetry in
craniofacial morphology 258, infections 112-113, 429-430 solders 230
260-264, 261, 262, 267-268 plague (7. pestis) 112 injuries
malocclusion 270-273 treponematosis 102 battle of Towton 162, 163
cultural influences 3 mitochondrial 365-368 decapitation 162-163
deficient see nutritional stress in the Americas 381-384, 382, farming-related 121, 130
elemental analysis 347-353 399, 400 rib fractures 117-118
iron 31-32, 352-353 postmarital residence 396, 397 lead levels 354, 355
isotope analysis 301-302, dolichocephalic head shape leprosy 111, 125
355-356 258-259 marine diet 319
and adult life history 330-332 domestic violence migration 343
bivariate 813C/815N analysis children 174, 177 oral health 82, 85-86
333-338, 334, 335 females 135-136, 157-158 osteoarthritis 180, 192, 203-204
8i3C values 302-304, 303 Donner Party pioneers 420 rickets 21-22
815N values 320-321, 321 double zone osteons 63 spondylolysis 120-121
diet-to-tissue spacing drought 173 Tudor navy personnel 192, 248
303- 304 Dublin (Viking period) 343-344 vertebral trabecular bone
maize consumption 304, stability 63-64
304- 316, 305, 309, 311, Easter Island (Rapa Nui) 155-156, weaning 327-329, 328
312 399-400 entheseal changes 206-212, 207
marine resources 304, 310, ebumation 181, 185 environmental stress see stress,
318-320, 319, 322-323, Ecuador 19, 205, 313 environmental
333-338 Efe tribe (Africa) 74-75, 78, 430 epidemiology, definitions 66
meat eating 313, 314, 321, 322, Egypt erosion (tooth) 277
331 expansion into Nubia 156 eruption of teeth 25, 76
millet consumption 316-318, familial groupings 390-391 Eskimo see Inuit
317 long bone morphology and the ethics 428-429
weaning 323-330, 325, 326, foraging-to-farming ethnogenesis, ethnogenetic 357,
328 transition 240-241, 241 387, 424
maize see maize see also Nile Valley Euroamericans
marine see marine diet elbow osteoarthritis 193-194, accidental injuries 118
meat see meat eating 202-203 bone histomorphology 246-247
millet 59-60, 316-318, 317 elemental analysis 347-348 craniofacial morphology 257, 258
and periodontal disease 81-83 barium 350-351,352 degenerative joint pathology 192
rice 71, 281 of collagen and apatite 355-356 diet 296-297, 315-316, 329
sex differences iron 352-353 growth rates 12-13, 13
caries 74-76 lead 353-354, 355 kinship 391
protein/cereal balance mercury 354-355 military personnel 166-168, 192,
313-316, 314, 317, multi-element analysis 352 211, 316, 354
330- 332, 332, 350-351 strontium 348-350, 350, 352 in native society 400-401
socioeconomic status zinc 352-353 see also isotope pioneer families 192, 315-316, 391
caries 77-78, 313 analysis Donner Party disaster 420
maize consumption 77-78, 306, Elephantine (Egypt) 390-391 stature 15-16
312, 313-315 enamel defects 44-57, 45 tooth wear 296-297
millet 316, 317 age at which stress occurred 49, Europe
protein/cereal balance 77-78, 50, 50-51 biodistance analysis 372-373
331- 332, 349 and caries 56-57 upper limb asymmetry
sorghum 318 duration/severity of stress 51-52 229-230, 248
strontium 348-350, 350, 352 hypocalcification 46, 56-57 bone remodeling due to
and tooth wear 277-288, 281, hypoplasia 44-48, 45, 47, 52-57, mechanical forces, lower
285, 286, 292, 292-297 56 limb 253
598 Index

Europe (conf.) diet Florida (USA), Spanish era


craniofacial morphology 260, 272 caries 74-76 biodistance analysis
dentition protein/cereal balance 74-75, kinship patterns 392, 393
caries 72, 77 313-316, 314, 317, temporal changes 361,
damaged teeth 298, 299 330-332, 332, 350-351 386-388, 387, 388
periodontitis 84-86 entheseal changes 210-211 bone remodeling due to
worn teeth 280, 284-287 fertility 178, 323 mechanical forces 235-237,
diet 72, 316, 317, 318-320, 319 growth/stature 20 236
foraging-to-farming transition injuries caries 70, 78
318-320, 319, 372-373 accidental 129 cranial porosities/anemia 37-38
fracture management 169 violent 134-137, 139, 141-143, diet 70, 294, 335, 335-336
growth rates 10, 13, 14, 24-25 157-158, 174 exploitation of indigenous
heavy metals 354-355, 355 osteoarthritis 188, 195-197, 203 population 164, 199-200
leprosy 109-111, 354-355 osteon density 247 mobility 251
osteoarthritis 196, 203-204 pelvic shape 21 osteoarthritis 199-200
paleodemography 405, 415 periosteal reactions 94-95 periosteal reactions 90-91, 95
plague 112 postmarital residence 345-346, population decline 415-418, 417
rickets 22, 23 392-396, 395 spondylolysis 121
spinal conditions 63-64, 121, 192 spondylolysis 121 treponematosis 101
squatting 204 tuberculosis 108 violent injuries 164, 165
stature 18, 20 see also pregnancy fluctuating asymmetry in dentition
treponematosis 99, 354-355 femur 28-29, 30
tuberculosis 103-104 bone remodeling due to folic acid 32
Viking migration 343-345, 345 mechanical forces 223-224, food processing
violent injuries 159-162, 163 225 cooking 260, 267-268, 280, 354
decapitation 160, 162-163, 175 age-related changes 241-246, maize
weaning 327-329, 328 244, 245 degenerative joint conditions
see also individual countries asymmetry 230-231 194, 205
exogamy 371 mobility 250-253, 251, 252, effect on nutritional value 16,
eye orbits 254 35
cribra orbitalia 33-34, 34, 39-41 size standardization 220 and tooth wear 279
scurvy 39, 40 temporal trends 231-241, 234, upper limb bone remodeling
236, 239, 247-249 228, 232-233, 238
facial injuries 173-174 diaphyseal bowing (rickets) 22-23 manioc 76, 290-291
farmers see agriculturalists; diaphyseal flattening 23-24, 249 and tooth wear 278-280,
foraging-to-farming length see stature 290-291, 296
transition neck-shaft angle 253-255 foot
females squatting 204 metatarsal atrophy in leprosy
bone histomorphometry 63 fertility 110
bone loss/osteoporosis 58-59, and breastfeeding/weaning 323 metatarso-phalangeal joint, and
125-126, 333 and female workload 178 kneeling 205, 206
bone remodeling due to in paleodemography 411-418 osteoarthritis 194, 199, 205
mechanical forces decline with agricultural foragers 178
foraging-to-farming transition intensification 414-418, bone remodeling due to
228-229, 232-233, 238, 417 mechanical forces 222-227,
240-241, 241 juvenility index in the 223, 225, 251
mobility 251, 252, 253 foraging-to-farming cranial porosities 36
and socioeconomic status 230 transition 413-414, 414, dentition
terrain 224 415, 416 caries 69-70, 70, 72-76
cranial porosities 38 population pyramids 411, 412 periodontal disease/tooth loss
death rates in disasters 420 fertilizers 320 81-83
dentition fiber processing (grooving of teeth) tooth wear 279, 284-287, 285,
caries 73-77 288-290, 289 286, 292-297
damaged teeth 298 Fiji 337-338 growth rates 12
enamel defects 57 firearms injuries 155, 164, 167-168 mobility 251, 252
periodontal disease 76, 83-85, fish/fishing see marine diet osteoarthritis 188-189, 190,
84 Fishergate House (London) 24-25, 193-196, 203
tooth wear 287-290 230 periosteal reactions 88-94
'

Index

sex differences 74-76,195-196,251 periodontal disease 80-81 handedness


violence among 175 stature 15 asymmetry of arm bones 214,
foraging-to-farming transition tooth size 26, 274 227-231, 231, 248
biodistance analysis 372-373 Georgia Bight (USA) asymmetry of osteoarthritis 193
bone remodeling due to mechanical caries 73 patterns of violent injuries 124,
forces 228-229, 231-241, 234, craniofacial morphology 263 140-141, 151, 162
236, 239, 247-249 diet 334, 334-336, 335 Harappan civilization see Indus
813C values and a marine diet humeral bilateral asymmetry Valley civilization
318-320, 319 228-229 Harris lines 42, 42-44
craniofacial adaptations related to mobility 251 Harvie family (Canadian pioneers)
masticatory behavior osteoarthritis 198-202, 201 315-316
260-264, 261, 262, 272-274 periosteal reactions 94 Haversian canals 60
decline in health/nutritional postmarital residence 393 Hawaii (USA) 298
status 10, 16-18, 424-425 St. Catherines Island mission 27, Hawikku (New Mexico) 166, 394
dentition 37-38, 235-237, 236, 334 head
asymmetry 28 temporal trends in bone morphology carrying loads on 195
enamel defects 53-54 233-237, 234, 236 see also craniofacial morphology;
malocclusion 272 tuberculosis 107-108 entries a t cranial
tooth size 26, 273-274 Germany 160-161 Heame, Samuel 145
tooth wear 279-282, 281 Giecz (Poland) 332, 332 heart disease 85
entheseal changes 210-211 gingivitis 76 heart, sacrificial removal 152
growth rates 10 Gotland (Sweden) 161-162, 169,210 hemolytic anemia 31, 33
osteoarthritis 197-203 Grasshopper Pueblo (Arizona) 211, herbivores, isotope and elemental
population increase 413-414 308, 350-351 analysis 320, 348, 350
strontium levels 349 Great Basin (USA) Herodotus 256
forearm biodistance analysis 383 hide chewing 84, 188, 287
fractures 122-124, 123, 126, 157 bone remodeling due to hip joint 179, 194, 204
Colies’ 122, 124 mechanical forces 222-225, histomorphometry of bone 60-64,
osteoarthritis 193-194, 202-203 223, 225, 251 61,62, 246-247
Fort Ancient sites (Ohio River valley) diet and maize consumption hominin species
306-307 308, 313 accidental injuries 127-129, 128
Fort William Flenry (New York) osteoarthritis 188-189 bone morphology and mechanical
166-167 tooth grooves due to fiber forces 231, 239, 239, 246
fractionation factor in stable isotope processing 288-290, 289 humeral bilateral asymmetry
analysis 303-304 Great Britain see England; Scotland 231
fractures see injuries Great Plains (USA) enamel hypoplasia 53
French and Indian War (Seven Years biodistance analysis 384-385, osteoarthritis 186-188
War) 166-167 398, 400-401 squatting 204
future research 429-432 bone remodeling due to tooth loss 81
mechanical forces 237-238 tooth size reduction 274
gastrointestinal infections 34, 36-38 diet 307-308 Honduras 57, 310, 312-314
gender differences see sex violent injuries 136-138, 167-168 hookworm 34, 37
differences see also Ankara people Hooton, Earnest A. 404
genetic variation in biodistance Great Salt Lake area (Utah) 225, Hoppa, Robert 403
analysis 359-360, 362, 308, 309 horseback riding 194, 238
367-368 Greece 180, 319, 405 humerus
Americas 380-382, 382 Greenland 267-268, 290, 346 damaged 128, 165
Great Basin area 383 growth rate indicative of cannibalism
Southwest 384 and environmental stress 9-14, 146-147, 147
East and Southeast Asia 377 64, 425-426 entheseal changes 207
Europe 372-373 Harris lines 42, 42-44 osteoarthritis 185, 193-194
postmarital residence 396, 397 skeletal compared with dental remodeling due to mechanical
genetic variation statistic (Fsp) 359-360 development 25 forces
genetics vertebral 24-25 bilateral asymmetry 214,
anemia 31, 33 Guatemala 9-10, 52-53 227-231, 231, 248
caries 68 gum disease see periodontal disease and environmental stress 223
handedness 227-228 gunshot wounds 155, 164, foragers 223-224
pathogens 429-430 167-168 kayaking 225-227
humerus (conf.) Indian subcontinent Inuit
maize processing 228, 232-233, biodistance analysis 373-374 bone remodeling due to
238 entheseal changes in the arm mechanical forces 225-227
temporal trends 228-229, 209 craniofacial morphology
231-241, 234, 236, 241 leprosy 110 260-261, 265-266
Hungary 77, 84-85 osteoarthritis of the neck 195 dentition 266, 287, 290
hunter-gatherers see foragers tooth loss 82 damaged teeth 297-298, 299
hunting tooth wear 284, 291, 296 entheseal changes in the arm
and diet 74 Indus Valley civilization 195, 284, 209-210
entheseal changes 209-210 291, 296 Harris lines 43
injuries 128-129 biodistance analysis 373-374 osteoarthritis 188, 191, 193, 196,
osteoarthritis ofthe elbow 193-194 industrial societies 268
hypermineralization zones in bone cranial porosities/anemia 35 osteon structure 63
63 osteoarthritis 179-180 spondylolysis 119
hypocalcification of dental enamel periodontal disease 82 violent injuries 143-146
46, 56-57 rib fractures 117-118 Irene Mound site (Georgia) 107-108,
hypoplasia of dental enamel 44-46, rickets 21 336, 393
45,47 spondylolysis 120 iron 31-32, 352-353
age at which stress occurred 49, weaning 327-329, 328 see also cranial porosities, and
50, 50-51 infant mortality 330, 405, 408 anemia
causes and associations 46-48, life table analysis 407 Isola Sacra cemetery (Roman
52-57, 56 infant weaning see weaning Empire) 51, 60, 327, 330-331
duration/severity of stress 51-52 infections 66, 113-114 isotope analysis 302
hypothesis testing 424-427 DNA analysis 102, 112-113, archaeological importance of
429-430 technique 301-302,
Illinois (USA) see Cahokia; Dickson epidemiology 66 338-339, 355-356
Mounds gastrointestinal (leading to bivariate 813C/815N analysis
Illinois River Valley (USA) anemia) 34, 36-38 333-338, 334, 335
biodistance analysis 385-386, and growth rates 9, 12 carbon see carbon (C), analysis of
389-390, 390, 396, 397 malaria 33, 112 stable isotope ratios
bone mass 14, 60 periostitis and osteomyelitis 12, collagen or apatite 302-304, 338,
craniofacial morphology 269 86, 86-96, 87 355-356
enamel hypoplasia 54 plague 112 and general health 330, 333
growth rates 10-11, 24, 44 and population health 423 nitrogen see nitrogen (N), analysis
infections 89-90, 100, 106-107 sex differences 94-95, 102, 108 of stable isotope ratios
osteoarthritis 196-197, 202 socioeconomic status 95-96, 102, oxygen 346-347
paleodemography 407, 413, 414 108, 111 strontium 339-346, 340, 344, 345
pelvic shape 21 see also leprosy; periodontal see also elemental analysis
social status 197 disease; treponematosis; Israel, antiquities law 428
stature 19 tuberculosis see also Natufian culture
temporal trends in bone injuries 115 Italy
morphology 231-233, 234 accidental see accidental injuries bone morphology and mobility
tooth wear 295 causes of increased fracture risk 253
violence 132-135, 174-175 125-127 caries 77
Inca Empire 154-155 cranial see cranial injuries diet 316, 332
incidence, definition 66 distinguishing antemortem from Isola Sacra cemetery (Roman
incisor teeth postmortem damage 116, Empire) 51, 60, 327, 330-331
chewing behavior 265 117 rickets 22
congenital syphilis 97, 100 sex differences see sex stature 20
damaged 298, 299 differences, injuries
grooved 288-290, 289 socioeconomic status 118, 142, Japan (and the Japanese)
lost 83-84 154 biodistance analysis 375-376
notching 290 tooth loss 84, 84 changes in craniofacial
shovel-shaped 266 treatment 168-171, 170 morphology 257, 272-273
wear patterns 287-288 violent see violence/violent diet 281, 337
lingual surface 290-292, 291 injuries enamel defects 54
Indian Knoll (Kentucky) 28, interpersonal violence see violence/ lead toxicity 354
195-196, 269, 405 violent injuries tooth wear 281, 288, 295
Index
V.

Jebel Sahaba (Sudan) 370 leprosy 108-111, 110, 125 Lucy (australopithecine) 186-187
Johnstown Flood (Pennsylvania, mercury treatment 354-355 lumbar spine 119, 195
1889) 420-421, 421 rhinomaxillary syndrome 109,
joints 109 Mac Bac (Vietnam) 72-73
anatomy 179 Levant Magdalenska Gora (Slovenia) 316
kneeling 205, 206 Bronze Age-Iron Age transition Mahalanobis D2 statistic 359
morphology and mechanical 368 Maitas-Chiribaya (Chile) 19, 96
loading 240 craniofacial morphology 264 maize
osteoarthritis see osteoarthritis dentition caries 70-71, 77-78
squatting 204 caries 71 consumption as charted by carbon
Jomon period (Japan) 51, 53, 272, tooth size reduction 273-274 isotope ratios
337, 375-376 tooth wear 280, 281, 295-296 bivariate 513C/516N analysis
Jordan 205, 391-392 endogamy 391-392 333-338, 334, 335
juveniles see children entheseal changes 210-211 comparison with marine diet
foraging-to-farming transition 333-336, 334, 335
kayaking 188, 205, 209, 225-227 210-211, 264, 273-274, 280, in Euroamericans 315-316
Kentucky (USA) 28, 195-196, 269, 349 in Mesoamerica 308-310, 312,
405 metatarso-phalangeal 312- 314
Kerma (Sudan) 156 modifications 205 in North America 304,
Khoi pastoralists 370-372 thalassemia 31, 33 304-308, 305, 309, 313
kinship patterns, biodistance see also ^atalhoyiik (Turkey) sex differences 313-315, 314
analysis 390-396, 393, 395, Libben site (Ohio) 12, 56, 127, 169 socioeconomic status 306, 312,
397 life expectancy 402 313- 315
kneeling 205, 206 see also age-at-death in South America 310-311,
Kodiak Island (Alaska) 145-146, life tables 406-409 311, 313
388-389 lingual surface attrition of the nutritional stress 16-18
Roger’s Island (Alabama) 138-139, maxillary anterior teeth and anemia 35, 37
272 290-292, 291 and bone remodelling 60-63
Kulubnarti (Sudan) 34-35, 156-157, linguistics, Great Basin area 383 and lower bone mass 59-60
243 Little Bighorn, battle (1876) 167-168 processing
Little Ice Age 346 degenerative joint conditions
La Chapelle-aux-Saints (France) loading forces on bone 215, 216 194, 205
187-188 London (UK) effect on nutritional value 16,
La Florida (Ecuador) 313 industrial era 35
La Florida (North America) osteoarthritis 180 and tooth wear 279
see Florida (USA), Spanish era rickets 21 upper limb bone remodeling
La Plata Valley (New Mexico) 95, spondylolysis 120 228, 232-233, 238
148 weaning 327-329, 328 protein/cereal balance 74-75,
La Real (Peru) 151 Medieval 24-25, 85-86, 230 313-316, 314, 330-332,
Lake Baikal region (Russia) 203, 210, lower limb 332, 350-351
282, 338 bone remodeling due to malaria 33, 112
Lamanai (Belize) 54-55, 309-310, mechanical forces 223-224, males
312, 314 225 bone remodeling due to
Lambayeque Valley (Peru) age-related changes 241-246, mechanical forces
colonial period 13, 37, 176-177, 244, 245 foraging-to-farming transition
201 asymmetry 230-231 228-229, 232-233, 240-241,
Sican culture 96, 152, 153, 197, inactivity 215 241
398-399, 400 mobility 250-253, 251, 252, mobility 223-224, 251, 252,
Lapa do Santo (Brazil) 149 254 253
Larson Village (South Dakota) temporal trends 231-241, 234, and socioeconomic status 230
137-138 236, 239, 247-249 death due to starvation 420
lane (megaliths) 121 diaphyseal bowing (rickets) 21-23 dentition
Lawson, John 101-102 diaphyseal flattening 23-24, 249 caries 73-77
lead (Pb) 353-354, 355 femoral neck-shaft angle 253-255 periodontal disease 83-85, 84
Leavenworth (South Dakota) 398 kneeling 205, 206 tooth wear 287-290
leg see lower limb periosteal reactions 88-89, 97, 98 diet
leguminous plants 320 squatting 204 caries 74-76
Lengua people (Paraguay) 29 sword cuts 161 entheseal changes 209-211
Index

males (conf.) Mayan people and osteoarthritis 232


growth/stature 20 colonial period burial 392 sex differences
injuries dentition foraging-to-fanning transition
accidental 129 enamel defects pre- and post­ 228-229, 232-233, 238,
tooth loss 84, 84 contact 54-55 240-241, 241
vertebral 121-122 socioeconomic status and oral mobility 223-224, 251, 252,
violent 136-141, 143, 157-158 health 77-78, 85 253
osteoarthritis 188, 194-197, 203 diet (isotope analysis) 309-310, and socioeconomic status in the
osteon density 247 312, 313-315, 332 Mesolithic 230
periosteal reactions 94-95 growth rates/stature 9-10, 18 socioeconomic status 230, 248
postmarital residence 345-346, Mbuti tribe (Africa) 74-75, 78, 430 soldiers 230, 248
392-396, 395 McCutchan-McLaughlin site terrain 225, 253
tuberculosis 108 (Oklahoma) 136 Medieval Europe
Malheur wetlands (Oregon) 189, mean measure of divergence (MMD) biodistance analysis 373
222-225, 223 statistic 359, 360 dentition
malocclusion 270-273 meat eating caries 77
Mancos Canyon site (Colorado) iron 31 damaged teeth 298, 299
146-147 nitrogen isotope ratios 313, 314, periodontitis 84-86
mandible 321, 322, 331 diet 77, 316, 317
size reduction 260-262, 271-272 sex differences 74-75 fracture management 169
temporomandibular joint socioeconomic status 77-78 growth rates 13, 24-25
osteoarthritis 188, 268 strontium levels 349 leprosy 109-111, 354-355
tori 266-268 tooth wear 294 osteoarthritis 196, 203-204
manioc 76, 290-291 mechanical loading and craniofacial plague 112
Mariana Islands 121, 205 adaptations 257 rickets 22, 23
marine diet and age 261, 268-270 spinal conditions 63-64, 121
barium and Ba/Sr ratios 351, 352 Arctic peoples 265-268 squatting 204
bivariate 813C/815N analysis malocclusion 270-273 treponematosis 99, 354-355
333-338, 334, 335 masticatory behavior 258, upper limb asymmetry in military
813C values 304, 310 260-268, 261, 262 personnel 230, 248
and the foraging-to-farming oral hard tissue tori 266-268, 267 Viking migration 343-345, 345
transition 318-320, 319 supraorbital torus 264-265 violent injuries 159-162, 163
caries 73 temporal trends 257-264, 258, decapitation 160, 162-163, 175
dietary shift in coastal California 261, 262, 267, 272-273 weaning 327, 328
19, 73, 282, 336-337 tooth size changes 271-276, 274 megalithic construction 121
8 '5N values 322-323 tooth wear 278-288,281,285, 286 megaloblastic anemia 32, 38
osteoarthritis 202-203 mechanical loading and long bone Mehrgarh (Indus River valley) 284,
marriage systems remodeling 214-215, 255, 291
endogamy 391-392 430 men see males
exogamy 371 age changes 241-246, 244, 245 mercury (Hg) 354-355
postmarital residence patterns agriculturalists 225,226, 228-229 Mesoamerica
345-346, 392-396, 395, 397 bilateral asymmetry 214, colonial period burial 392
Mary Rose (Tudor warship) 192, 248 227- 231, 231, 248 dentition
masticatory behavior cross-sectional geometry caries 57, 77-78
craniofacial adaptations 258, 215-222, 218, 219 enamel defects 52-55
260-268, 261, 262 as applied to specific groups socioeconomic status and oral
malocclusion 270-273 222-246 health 77-78, 85
reduction in tooth size external measurements 247-255 diet (isotope analysis) 308-310,
271-276, 274 femoral neck-shaft angle 253-255 312, 313-315, 332
temporomandibular joint foragers 223, 251, 252 growth rates/stature 9-10, 18-19
osteoarthritis 188, 268 foraging-to-farming transition Monte Alban society 397-398
tooth wear 278-288, 281, 285, 228- 229, 231-241, 234, Mesolithic Age
286, 292, 292-297 236, 239, 241, 247-249 bone morphology and
see also dentition, histomorphometry 60-64, 61, 62, socioeconomic status 230,
extramasticatory uses 246-247 248
maxilla loading forces 215, 216 caries 72
infections 109, 109, 111 mobile lifestyles 250-253, 251, dietary shift compared with
size reduction 271-272 252, 254, 254 Neolithic 318-320, 319
Index 603

tooth loss 82-83 mutilation trophic level in food chain


tooth wear 280, 284 Arctic/Subarctic area 145 320-321, 321
violence 159-161 Europe 160, 162-163 weaning 323-330, 325, 326,
metatarsals North America 133, 133, 328
leprosy 110 137-138, 143, 167-168, 175 Nordin’s Index see cortical bone
osteoarthritis 194, 199, 205 South America 148-154, 150, mass
metatarso-phalangeal joint, and 151, 153 Norris Farms (Illinois) 132-135, 133,
kneeling 205, 206 Mycobacterial infections 134, 174-175
Mexico see leprosy; tuberculosis Norse people
agricultural intensification 92 mandibular tori 267-268
decreased mobility 252-253 Nasca (Peru) 152 migration 343-345, 345
diet 308, 349 Native American Graves Protection see also Scandinavia
Monte Alban society 397-398 and Repatriation Act (USA, North America
Teotihuacan society 1990 and 2010) 428-429 biodistance analysis
residence patterns 342-343, Native Americans see Mesoamerica; ethnic boundaries 389-390,
346-347, 394 North America; South 390
ritualized violence 152-154 America Euroamericans in native society
Michigan (USA) 135-136 Natufian culture 400-401
micronutrients caries 71 kinship and marriage 391-392,
iron 31-32, 352-353 tooth wear 280, 281, 295-296 393, 394, 395, 396, 397
zinc 352-353 upper limb use 210-211 linking living groups to
see also entries at vitamin Neandertal man ancestors 388-389
Middle East see Levant accidental injuries 127-129, 128 migration into and across the
migration bone morphology and mechanical continent 377-383, 382
biodistance analysis forces 231, 239, 239, 246 Spanish Florida 361, 386-388,
see biodistance analysis enamel hypoplasia 53 387, 388, 392, 393
isotope analysis 338-347, 340, osteoarthritis 187-188 within/into geographic regions
344, 345 squatting 204 383-386
milk teeth see deciduous teeth neck bone remodeling due to
millet 59-60, 316-318, 317 injuries 121-122, 127-129 mechanical forces
mineralization of bone 63, 216 osteoarthritis 195-196 agriculturalists 225, 226,
Minusinsk Basin (Russia) 317-318 Neolithic Age 228-229
Mississippi River valley (USA) 38, biodistance analysis 372-373 Aleutian islanders 225-227
96, 305-307 caries 71 Great Basin foragers 222-225,
Missouri River valley (USA) dietary shift compared with 223, 225, 251
see Ankara people Mesolithic 318-320, 319 and mobility 251-253, 252
mitochondrial DNA (mtDNA) 365-368 male mobility 253 temporal trends 228-229,
in the Americas 381-384, 382, population increase 413-414 231-239, 234, 236, 247-248
399, 400 tooth wear 280, 295 craniofacial morphology 258,
postmarital residence 396, 397 violence 160-161 261-265, 269
mobility weaning 325, 326 dentition
bone morphology 223-224, see also f atalhoyiik (Turkey) asymmetry 28
250-253, 251, 252, 254, neonates, Wilson band formation 48 caries 56-57, 69-71, 70, 73-74,
430 neural canal size 24-25 78
isotope analysis 338-347, 340, New Mexico (USA) 95, 148, 166, damaged teeth 298-300
344, 345 238-239, 394 enamel defects 51-57
Moche culture (Peru) 311, 311 see also Pecos Pueblo periodontal disease/tooth loss
molars Nile Valley see Nubia 81, 85
damaged 297-298, 299 nitrogen (N), analysis of stable tooth size 26-28
wom 68-69 isotope ratios tooth wear 282-283, 287-290,
see also caries adult life history 330-332, 332, 289, 294-295
Monte Alban (Mexico) 397-398 334 diet
mortality profiles 404-406 general health 330, 333 maize see maize
and life tables 406-409 marine diet 322-323 marine or terrestrial 19,
Moundville (Alabama) 95-96, 107, bivariate 813C/515N analysis 333-337, 334, 335, 351, 352
129-130, 140, 336 333-338, 334, 335 meat eating 313, 314, 349
multivariate analysis in biodistance meat eating 313, 314, 321, 322, sex differences 313, 314,
analysis 359 331 331-332, 350-351
604 Index

North America (cont) horseback riding 194 stature 14-20, 64-65


entheseal changes in the arm 211 sex differences 188, 194-196, weaning 12, 14, 51, 330
environmental stress 203
cranial porotic lesions 35-39 socioeconomic status 197 Oaxaca valley (Mexico) 92, 397-398
decreased bone mass 14, 59-60 temporal trends 198-203, 201 occlusive abnormalities 270-273
dentition 26-28, 51, 53-55 vertebral 194, 196 Ofnet (Germany) 160-161
growth rates 10-13, 11, 13, weapon use 193-194 Ohio (USA), Libben site 12, 56, 127,
43-44 paleodemography 405 169
histomorphometiy 60-63 disaster events 420-421, 421 Ohio River valley (USA)
rickets 22-23 infant mortality 405, 407 biodistance analysis 386
stature 16-20 juvenility index 413, 414, 416 craniofacial robusticity 263
susceptibility to violent attack Puebloan people 404 maize consumption 305-307
174-175 periosteal reactions 89-96 Oklahoma (USA) 136
injuries 172-173 repatriation of remains 388, omnivores 321, 348
18th century and 19th century 428-429 Ontario (Canada)
fighting 166-168 spondylolysis 121 diet
Arctic/Subarctic 143-146 treponematosis 97-102 19th century 315-316, 329
European conquest 163-166, tuberculosis 106-108 prehistoric 304-305, 305, 349,
165 nose 350
Great Plains 136-138, 167-168 leprosy 109, 109 growth rates 12-13
Midwest 127, 132-136, syphilis 97 rib fractures 118
174-175 Nubia Snake Hill military cemetery 167,
mutilation 133, 133, 137-138, anemia 33-35 192, 211, 316, 354
143, 167-168, 175 biodistance analysis of population Wise cemetery (biodistance
Pacific coast 123, 140-143, 144 structure over time 369-370 analysis) 391
and shift to agriculture bone histomorphometry 63 Oregon (USA) 189, 222-225, 223
129-130 craniofacial morphology osteitis 87
Southeast 130, 138-140, 164, 259-260, 261, 262 osteoarthritis 6, 212
172 dental asymmetry 29, 30 and age 180-181, 196-197
Southwest 146-148, 166 diet 318 comparative studies 181, 189-192
treatment 169 growth rates 14, 25, 243 and different kinds of physical
lead levels 354 injuries 124, 156-157 activity 179-180, 186-195
linguistics 383 treatment 169 farmers 180, 192, 194-196
missionization in Spanish Florida long bone morphology and the foragers 188-189, 193-196
biodistance analysis 361, foraging-to-farming foraging-to-farming transition
386-388, 387, 388, 392, 393 transition 240-241, 241 197-203
bone remodeling due to osteoporosis 59, 333 horseback riding 194
mechanical forces 235-237, periosteal reactions 92-93 weapon use 193-194
236 tooth loss 82-83 interpretation of data 181-186
caries 70, 78 tooth wear 281-282, 284 and long bone geometry 232
cranial porosities/anemia 37-38 Numic languages 383 in other hominin species 186-188
decreased mobility 251 nutrition see diet pathology 179, 181, 182
diet 70, 294, 335, 335-336 nutritional stress ebumation 181, 185
exploitation of indigenous bone mass 13-14, 58-60, 223 spinal 181, 186, 187
population 164, 199-200 cranial base height 20-21 sex differences 188, 194-196, 203
osteoarthritis 199-200 cranial porosities 30-41, 32, 34 socioeconomic status 197
periosteal reactions 90-91, 95 dentition and spondylolysis 120
population decline 415-418, 417 enamel defects 46, 47 temporal trends 197-204, 201
spondylolysis 121 eruption rate/timing 25 temporomandibular joint 188, 268
non-indigenous racial groups tooth size 26, 49, 273 vertebral 181, 186, 187, 192,
see African-Americans; elemental analysis 194-196
Euroamericans iron 352-353 osteomalacia 21-23
osteoarthritis zinc 353 osteomyelitis 87, 87-88
and age 196-197 growth rates 9-14, 64 see also periostitis/periosteal
comparative studies 189-192 vertebral 24-25 reactions
food processing 194 and infections 9, 91 osteons, secondary 60-64, 246-247
in foragers in marginal settings pelvic morphology 21 osteophytes 179, 181
188-189, 190 rickets 21-23, 23 osteoporosis 57-60, 125-126, 333
Index 605

Oxus culture (central Asia) 374 diet 310-311, 311, 336 postmarital residence 394,
oxygen (0) isotopes 346-347 kinship in a Sican burial 398-399, 395
400 bone histomorphometry
Pacatnamu (Peru) 150, 150 osteoarthritis 193, 197, 201 246-247
Pacbitun (Belize) 310, 312, 312 periosteal reactions 95-96 cranial porosities 35
Pacific islands see Easter Island; Fiji; socioeconomic status 47-48, 78, 96, diet 63, 308, 350-351
Flawaii; Marianas Islands 197 entheseal changes in the arm
paleodemography 402-404, 421 Tiwanaku state 341-342 211
disaster events 420-421, 421 trephination 169-170, 170 osteoarthritis 191
and fertility/birth rate 411-418 tuberculosis 104-105, 107-108 paleodemography 404
decline with agricultural violent injuries 123, 150, violent injuries 146-148, 166
intensification 414-418, 417 150-152, 151, 153, 154-155 pygmy forager tribes (Aka, Mbuti,
juvenility index in the Pete Klunk mound (Illinois) 197, Efe) 74-75, 78, 430
foraging-to-farming 396, 397, 407
transition 413-414, 414, pewter 354 racial groups, craniofacial
415, 416, 416 photosynthetic pathways morphology not a good
population pyramids 411, 412 (813C values) 302-303, 303 guide 257-259
infant mortality 405, 407-408 Phum Snay (Cambodia) 159 radius
life table analysis 406-409 physical activity Colles’ fractures 122, 124
mortality profiles 404-406 and bone mass 59 entheseal changes 207
problems entheseal changes 206-212, 207 Rapa Nui (Easter Island) 155-156,
age estimation 407, 409-410, and female fertility 178 399-400
418-419, 419 foragers 178 reduction of fractures 169
excavation techniques 408-409 osteoarthritis see osteoarthritis relatedness, analysis see biodistance
underrepresentation of and remodeling of bone analysis
juveniles 407-408, 413 see biomechanics and remodeling of bone
variability in preservation of skeletal morphology due to mechanical forces
skeletal elements 403 rib fractures 117-118 see biomechanics and
timing of weaning 323, 330, 408 spondylolysis 118-122, 119 skeletal morphology
Panama 309 phytates 31, 34-35 histomorphometry 60-64, 61, 62,
Paraguay 29 Pickwick Basin (Alabama) 193, 207, 246-247
parasitic infestations 34, 37 229, 231-233, 272 repatriation of human remains
parry fractures of the forearm pinta 96-97 388-389, 428-429
122-124, 123 plague 112 residential change 338-347, 340,
pastoralism 54, 253, 370-372 Point Hope (Alaska) 260-261 344, 345
Patagonia (South America) 266 Poland 332, 332 rhinomaxillary syndrome (leprosy)
Pecos Pueblo (New Mexico) 166, poor house inhabitants 248 109, 109
191, 246-247, 308, 404 population density rib lesions
pelvic shape/deformity 21 and spread of disease 89-90,93-94 fractures 117-118, 118
percent cortical area index (PCCA) and violence 172-173 tuberculosis 103, 105, 107
219, 249 see also urbanization rice 71, 281
perikymata 45, 47, 52 population pyramids 411,412 rickets 21-23, 23
periodontal disease and lost teeth porotic hyperostosis 30-41, 32 riding 194, 238
78-81, 79, 82 Portugal 72, 280, 284-287, 320 rigidity of bone 218-220, 219
in pregnancy 76 posture ritualized violence 148-154, 150,
sex differences 83-85, 84 kneeling 205, 206 153, 177
socioeconomic status 84-85 squatting 204 Riviere aux Vase (Michigan)
and systemic health 85-86 Pott’s disease 103, 105 135-136
temporal trends in prevalence pregnancy Rodeo athletes 128
81-83 dentition 76-77 Roman Empire
periostitis/periosteal reactions 12, nitrogen balance 333 decapitation 162
86, 86-96 prevalence, definition 66 diet 330-331
Peru principal component analysis 359 fractures 60, 169
colonial period projectile injuries lead levels 354, 355
environmental stress 13, 37 arrows (orspears) 134,142-143,160 migration 345
osteoarthritis 201 firearms 155, 164, 167-168 weaning 51, 325-327
periosteal reactions 95 Pueblo Bonito (Florida) 394, 395 Ruffs mobility index 250-253, 251,
structural violence 176-177 Puebloan societies 252, 254
dentition 28, 47-48, 78, 82 biodistance analysis 384 Russia see Siberia
606 Index

saber-shin 97, 98 second moments of area 218-220, shovel incisors 266


Sacred Ridge (Colorado) 148 219 Siberia
sacrificial rituals 148-154, 150, sedentary lifestyle 59, 89-90, 93-94, diet 282, 317-318, 338
153, 177 192, 248-249 entheseal changes 210
Sadlermiut Inuit (Canada) 188 selectivity in the archaeological origin of Native Americans
Sahara desert, pastoralism 54 record 422-423 377-383
St. Catherines Island (Georgia, Santa Seven Years War (French and Indian osteoarthritis 203
Catalina de Guale) 27, 37-38, War) 166-167 Sican culture (Peru) 96, 152, 153,
235-237, 236, 334 sex differences 197, 398-399, 400
St. Jorgensgard (Denmark) 109 bone loss 58-59, 125-126, 333 sickle cell anemia 33
San Cristobal (New Mexico) 166 bone remodeling due to Sigtuna (Sweden) 332
San foragers 371 mechanical forces skull see craniofacial morphology;
San Luis de Apalache (Florida) 294, foraging-to-farming transition entries at cranial
392 228-229, 232-233, 238, Slovenia 316
San Pedro de Atacama (Chile) 154 240-241, 241 Snake Hill military cemetery
San Pedro de Morrope (Peru) mobility 223-224, 251, 252, 253 (Ontario) 167, 192, 211, 316,
176-177 and socioeconomic status in the 354
San Pedro y San Pablo de Patale Mesolithic 230 socioeconomic status
(Florida) 392, 393 cranial porosities 38 biodistance analysis 397-399, 400
sanitation, poor 16, 36-38 death rates in disasters 420-421, bone morphology 230, 248
Santa Barbara Channel Islands 421 cranial porosities 38-39
(California) dentition dentition
cranial porosities 36 caries 73-77 caries 77-78, 313
dentition 55, 73, 282 damaged teeth 298 enamel defects 47-48, 52,
dietary shift 19, 73, 282, 336-337 enamel defects 57 55-56, 56
environmental stress 19, 36, 55 periodontal disease/tooth loss tooth loss 84-85
osteoarthritis 202-203 76, 83-85, 84 tooth wear 288, 296-297
periosteal reactions 93-95 tooth wear 287-290 diet
stature 19 diet and caries 77, 313
violent injuries 123, 140-143, caries 74-76 maize consumption 77-78, 306,
144 protein/cereal balance 74-75, 312, 313-315
Santa Catalina de Guale (Georgia, 313-316, 314, 317, millet 316, 317
St. Catherines Island) 27, 330-332, 332, 350-351 protein/cereal balance 77-78,
37-38, 235-237, 236, 334 entheseal changes 209-211 331-332, 349
Santa Catalina de Guale de Santa growth/stature 20 entheseal changes 211
Maria (Florida, Amelia injuries growth rates 9-10, 14
Island) 70, 235-237, 236, accidental 129 injuries
303 tooth loss 84, 84 accidental 118
Saunaktuk (Canada) 144-145 vertebral 121-122 violent 142, 154
scalping 133, 133, 137-138, 175 violent 134-143, 139, lead toxicity 354
Scandinavia 157-158, 174 leprosy 111
dentition 77, 280 osteoarthritis 188, 194-196, 203 osteoarthritis 197
diet 280, 318,319, 332 osteon density 247 pelvic shape 21
entheseal changes 210 periosteal reactions 94-95 periosteal reactions 95-96
leprosy 109-110 postmarital residence patterns rickets 22
mandibular tori 267-268 345-346, 392-396, 395, 397 stature 19-20
mercury 354-355 treponematosis 102 structural violence 176-177
migration 343-345, 345 tuberculosis 108 treponematosis 102
osteoarthritis 196 shamans 19, 96 tuberculosis 108
treatment of injuries 169 Shanidar Cave (Iraq) 127 soldiers
violent injuries 159-162, 169 sharp force trauma humeral remodeling 230, 248
Schmorl’s depressions/nodes 187, projectile injuries 134, 142-143, osteoarthritis 192
192 160 violent injuries 161-162, 163,
sciatic notch width 21 sword injuries 161, 163, 164 166-168
Scotland 343 short tandem repeats 365 see also archers
scraper use 211 shoulder sorghum 318
scurvy 39, 40 osteoarthritis 188, 196 South Africa 264, 287, 323,
seafood see marine diet severed 164, 165 370-372, 428
Index 607

South America dentition Taforalt (Morocco) 298


biodistance analysis 382-383, asymmetry 28-29, 30 Talheim (Germany) 160
398-399, 400 development rate 25 tannates 31
colonial period enamel defects 44-57, 45, Tarim Basin mummies (China) 375
environmental stress 13, 37 47, 48, 49 Tatham Mound (Florida) 101, 164
osteoarthritis 201 tooth size 26-28, 49, 273 teeth see dentition
periosteal reactions 95 developmental origins of disease temporomandibular joint
structural violence 176-177 7-8, 54, 64, 431 osteoarthritis 188, 268
cranial porosities 37 diaphyseal deformation 22-24, 23 size reduction 263
craniofacial morphology 263, 266 and elemental analysis Tennessee (USA) 94, 130, 195-196
dentition iron 352-353 tennis players 214, 243
asymmetry 28-29 zinc 353 Teotihuacan (Mexico) 152-154,
caries 75-76, 78 growth rates 9-14, 24-25, 64, 342-343, 346-347, 394
tooth loss 82, 83 425-426 terrain, femoral morphology 225,
tooth wear due to manioc Harris lines 42, 42-44 253
preparation 290-291 and infections 9, 91 tetracycline 92-93, 93
diet 16-18, 310-311, 311, 313, and isotope analysis 330, 333 Thailand 130, 345-346, 377
336 pelvic morphology 21 thalassemia 31, 33
growth rates 13 and prevalence of violent injuries thoracic spine 104, 121-122, 192,
joint changes associated with 154, 166, 172-175 194
kneeling 205 skeletal and dental development throwing, changes in the arm
migration 340-343 compared 25 193-194, 209-211
osteoarthritis 193, 197, 201 stature 14-20, 64-65 tibia
periosteal reactions 95-96 vertebral indicators 24-25, 63-64 diaphyseal shape 23-24, 250
spondylolysis 119 weaning 12, 14, 51, 330 periosteal reactions 88-89, 97, 98
stature 19 stress, mechanical 215, 217, 218, squatting 204
trephination 169-170, 170 219-220 sword cuts 161
tuberculosis 104-108 striae of Retzius 45, 47, 52 Tierra del Fuego (South America)
violent injuries 123, 148-155, Wilson bands 48, 48-49, 51, 266
150, 151, 153 53-55 Timucua people (Florida) 235-237
South Dakota (USA) 136-138, strontium (Sr) Tipu (Belize) 54-55, 392
174-175, 260-261, 398, elemental analysis (diet) 348-350, Tiwanaku state (South America)
400-401, 426 350, 352 341-342
spears 193-194, 209-211 isotope analysis (residence and Tombigbee River valley (Alabama)
see also projectile injuries, arrows migration) 339-346, 340, 140, 172
(or spears) 344, 345 Tombos (Sudan) 156
spine see vertebrae structural violence 176-177 tongue excision 137
Spitalfields (London) 120, 180, sucrose 70 torsional loading/stress 215, 217,
327-329, 328 Sudan see Nubia 218, 219-220
spondylolisthesis 121 Sully (South Dakota) 260-261, 426 torsional rigidity (3) 220, 223-224,
spondylolysis 118-121, 119 supraorbital torus 264-265 225
squatting 204 swaddling 22 Towton, battle (1461) 162, 163, 230
starvation cannibalism 147-148 Swan Creek (South Dakota) toxicity
stature, adult 14-20, 64-65 400-401 lead 353-354, 355
and growth rates 9-13, 64 Sweden 77, 196, 210, 332 mercury 354-355
Stillwater Marsh (Nevada) 189, battle of Wisby 161-162, 169 trabecular bone 63-64
222-225, 223, 383 swords trace element analysis 347-348
strength of bone 221 humeral changes due to sword use barium 350-351, 352
Streptococcus mutans 429-430 230 collagen or apatite 355-356
stress, environmental 8-9, 64-65 injuries caused by 161, 163, 164 iron 352-353
biocultural model 7, 8 syphilis 88-89, 96-98 multi-element analysis 352
bone histomorphometiy and congenital 97, 100 strontium 348-350, 350, 352
remodeling 60-64, 61, 62 cranial vault 97, 99, 101, 101 zinc 352-353
bone mass 13-14, 57-60, 58, endemic 96-97 trade networks, and spread of
63-64, 223 mercury treatment 354-355 disease 89
cranial base height 20-21 tibia 97, 98 trauma see injuries
cranial porosities (anemia-related) Syria (Abu Hureyra) 205, 280, Trelleborg (Denmark) 344-345, 345
30-41, 32, 34 295 trephination 169-170, 170
608 Index

treponematosis 88-89, 96-102 stature reduction 16 ritualized 148-154, 150,


congenital syphilis 97, 100 Utah (USA) 225, 308, 309 153, 177
cranial vault 97, 99, 101, 101 sex differences 134-143, 139,
mercury treatment 354-355 venereal disease (syphilis) 97, 99, 157-158, 174
tibia 97, 98 100, 354-355 socioeconomic status 142, 154
Trino Vercellese (Italy) 20, 77, 316, vertebrae South America 123, 148-155,
317 growth 24-25 150, 151, 153
trophic level of organisms osteoarthritis 181, 186, 187, 192, structural 176-177
and barium 350 194-196 treatment 168-171
and nitrogen isotopes 320-321, spondylolisthesis 121 trephination 169-170, 170
321 spondylolysis 118-121, 119 vitamin B3 (niacin) 16
and strontium 348 trabecular bone 63-64 vitamin B12 32, 38
tuberculosis 60, 102-108, 114 tuberculosis 103, 104, 105 vitamin C (ascorbic acid) 31
rib lesions 103, 105, 107 Vietnam 72-73 scurvy 39, 40
vertebral lesions 103, 104, 105 Vikings vitamin D 21-23, 23
Turkey, leprosy 110 migration 343-345, 345
see also fatalhoyiik (Turkey) violence 159 Wadi Haifa (Sudan) 33-34, 92-93,
violence/violent injuries 115, 156, 369
ulna 130-132, 132 warfare see violence/violent
entheseal changes 209 arrows (or spears) 134, 142-143, injuries
parry fractures 122-124, 123 160 Wari empire (Peru) 123, 151-152
undemutrition see nutritional stress Australian Aborigines 124, water-borne infections 34, 36-38
United Kingdom (UK) see England; 157-158 weaning
Scotland Cambodia 159 causes environmental stress 12,
United States of America (USA) cannibalism 145-148, 147, 149, 14, 51, 330
see North America 155-156, 173 timing of
upper limb causes and purpose of violent demographic effect 323
bone remodeling due to behavior 172-177 elemental analysis (Ba/Sr)
mechanical forces child abuse 174, 177 349-351
bilateral asymmetry 214, cranial see cranial injuries nitrogen isotope ratios
227-231, 231, 248 Easter Island 155-156 323-330, 325, 326, 328
and environmental stress 223 Europe 159-162, 163 Westerhus (Sweden) 77, 196
foragers 223-224 decapitation 160, 162-163, 175 Wharram Percy (England) 63-64,
kayaking 225-227 firearms 155, 164, 167-168 121, 327, 328
maize processing 228, 232-233, forearm fractures (may be wickerwork 289-290
238 accidental) 122-124, 123, Wilson bands 48, 48-49, 51,
temporal trends 228-229, 124, 157 53-55
231-241, 234, 236, 241 mutilation see mutilation Wisby, battle (1361) 161-162,
damage indicative of cannibalism North America 172-173 169
146-147, 147 18th century and 19th century Wise family (Canadian pioneers)
entheseal changes 206-212, 207 fighting 166-168 391
injuries Arctic/Subarctic 143-146 witches 148
accidental 126, 128 European conquest 163-166, Wolff, Julius 214
Codes' fractures 122, 124 165 Womble analysis 389, 390
parry fractures 122-124, 123 Great Plains 136-138, 167-168 women see females
violent 157, 164, 165 Midwest 132-136, 133,
osteoarthritis 185, 193-194, 174-175 Y chromosome lineages 381
202-203 Pacific coast 123, 140-143, 144 yaws 96-97
urbanization Southeast 130, 138-140, 164, Yersinia pestis (plague) 112
infections 103-104, 111 172
injuries 117-118, 129, 156 Southwest 146-148, 166 Zalavar (Hungary) 77, 84-85
osteoarthritis 191-192 Nubia/Nile Valley 124, 156-157 zinc (Zn) 352-353
(|BEA CAMBRIDGE STUDIES IN BIOLOGICAL AND EVOLUTIONARY ANTHROPOLOGY

This updated and revised edition of Larsen’s classic text provides a comprehensive overview of the
fundamentals of bioarchaeology. Reflecting the enormous advances made in the field over the past 20
years, Larsen examines how this discipline has matured and evolved in fundamental ways.

New to this edition:


The latest developments in stable isotope analysis reconstruction and the interpretation of social
organization, migration, warfare, ritual violence, and diet
Recent advances in demographic reconstruction in interpreting health, population structure, and
dynamics
• A greater geographic coverage, including the great expansion of bioarchaeology in Europe,
Southeast Asia, South America, and the Pacific.

The book underscores, for students and professionals alike, the central role of human remains in
interpreting life events and conditions of past and modern cultures. Supplementary material is available
online at www.cambridge.org/Larsen.

an impressive synthesis of new methodology ... a must read for anyone who seeks a
comprehensive overview of this dynamic field.”
Doug Ubelaker, Forensic A nthropologist

“...a comprehensive overview of how bioarchaeology enables us to understand the nuances of


our ancestors’ lives and deaths, and who we are today ... Larsen has again very successfully
synthesised a huge amount of information to provide an accessible te x t... I am convinced that this
book will again be welcomed by the bioarchaeological community.”
Charlotte Roberts, Durham University, UK

“As a handbook of bioarchaeological methods, there is nothing else like this on the market. This
book deserves a wide readership among those interested in learning about the analytical potential
of human remains.”
Debra Martin, University o f Nevada, Las Vegas, USA

"... [a] comprehensive, authoritative, empirically rich, and heavily referenced volume that is
simultaneously deliberately provocative and forward looking.”
George R. Milner, Penn State University, USA

“... Larsen captures not only the state-of-the-art of the discipline, but also its sheer breadth and
depth in this extraordinary synthesis. Bioarchaeology has been and will continue to be the go-to
volume in the field!”
Patricia Lambert, Utah State University, USA

Cover illustration: © The British Library Board. Royal 6 E. VI f.301. Detail of a


historiated initial 'C’(lericus): Clerics with leprosy receiving instruction from a bishop
C ambridge
UNIVERSITY PRESS
www.cambridge.org
ISBN 978-0-521-54748-2

Cover designed by Hart McLeod Ltd 9 780521 547482 >

You might also like