Th eDiTion
Th
eDiTion
offers a beautifully illustrated, full-color, site-byThe sixth edition of site survey of prehistory. Through five successful editions, instructors and students alike have been captivated by the book’s rich visual presentation, research, interpretations, and theories from the field. The new edition has been revised in accordance with both new discoveries in archaeology and the interests of readers.
gary Feinman
T. Douglas Price Features of the Sixth Edition The newest discoveries, research, and interpretations in archaeology have been added to the sixth edition. The visual program of the book has been further improved upon with new photos and line drawings throughout.
The first half of the book includes: new information on the relationship between archaeology and anthropology and historical archaeology (Ch. 1); a new section on argon-argon dating and new information on Atapuerca and the first Europeans (Ch. 2); a new hypothesis on meteors and Pleistocene extinctions (Ch. 3); new dates and a discussion of early, fully modern humans from Pinnacle Point, SA and a new discussion of first Americans (Ch. 4); new material on nitrogen isotopes in relation to the study of the paleo diet (Ch. 5); and new information on the domestication of plants and animals from Peru, Near East, China, Africa, and Europe (Ch. 6).
Price Feinman
ISBN 978-0-07-353105-2 MHID 0-07-353105-7 9 0 0 0 0
9
780073 531052 www.mhhe.com
MD DALIM 1049758 09/25/09 CYAN MAG YELO BLACK
New material in the second half of the book includes: new findings concerning the introduction of corn into the American Southwest (Ch. 7); early Mayan writing and new perspectives on the competition between different Mayan states during the Classic period (Ch. 8); new information on the mapping and size of Angkor (Ch. 10); the latest information on the Iceman, including the use of strontium isotopes to determine his place of origin, and exciting new data from the Stonehenge Riverside project, including interpretation of life and death contexts (Ch. 11). Finally, Chapter 12 has been expanded with additional information on career choices.
pri31057_fm.ppi-xx.qxd
10/15/09
11:10 AM
Page i
Images of the Past S I X T H
E D I T I O N
T. Douglas Price University of Wisconsin–Madison
Gary M. Feinman The Field Museum
< ( : < V : y < p : W
_
pri31057_fm.ppi-xx.qxd
10/15/09
11:11 AM
Page ii
IMAGES OF THE PAST Published by McGraw-Hill, an imprint of The McGraw-Hill Companies, Inc., 1221 Avenue of the Americas, New York, NY 10020. Copyright © 2010, 2008, 2005, 2001, 1997, 1993. All rights reserved. No part of this publication may be reproduced or distributed in any form or by any means, or stored in a database or retrieval system, without the prior written consent of The McGraw-Hill Companies, Inc., including, but not limited to, in any network or other electronic storage or transmission, or broadcast for distance learning. This book is printed on acid-free paper. 1 2 3 4 5 6 7 8 9 0 CCI/CCI 0 9 ISBN: 978-0-07-353105-2 MHID: 0-07-353105-7 Editor in Chief: Michael Ryan Sponsoring Editor: Gina Boedeker Marketing Manager: Pam Cooper Managing Editor: Nicole Bridge Developmental Editor: Janice Wiggins-Clarke Production Editor: David Blatty Manuscript Editor: Thomas L. Briggs Design Manager: Preston Thomas Cover Designer: Mary-Presley Adams Photo Research: Brian Pecko Production Supervisor: Tandra Jorgensen Composition: 9.25/12 Palatino by Thompson Type Printing: 45# Pub Matte, Courier Cover Image: Mexico, Tabasco, Parque-Museo La Venta, Olmec carved head. The Olmecs were of the first civilizations in Mexico and are renowned for their giant carvings of heads, some of which are up to 3 metres tall. © Robert Frerck/Stone/Getty Images. Credits: The credits section for this book begins on page C-1 and is considered an extension of the copyright page. Library of Congress Cataloging-in-Publication Data Price, T. Douglas (Theron Douglas) Images of the Past / T. Douglas Price, Gary M. Feinman.—6th ed. p. cm. Includes bibliographical references and index. ISBN-13: 978-0-07-353105-2 ISBN-10: 0-07-353105-7 1. Prehistoric peoples. 2. Antiquities, Prehistoric. 3. Archaeology. 4. Indians—Antiquities. I. Feinman, Gary M. II. Title. GN740.P75 2010 930.1—dc22
_ _
2009040460
The Internet addresses listed in the text were accurate at the time of publication. The inclusion of a Web site does not indicate an endorsement by the authors or McGraw-Hill, and McGraw-Hill does not guarantee the accuracy of the information presented at these sites. www.mhhe.com
pri31057_fm.ppi-xx.qxd
10/2/09
8:03 AM
Page iii
For Anne Birgitte Gebauer and Linda Nicholas
pri31057_fm.ppi-xx.qxd
10/2/09
8:03 AM
Page iv
<
Brief Table of Contents PREFACE ABOUT
xi
THE
CHAPTER
AUTHORS
xvii
ONE
Principles of Archaeology 1 CHAPTER
TWO
The First Humans 35 CHAPTER
THREE
Out of Africa: Homo Erectus 73 CHAPTER
FOUR
The Hunters 105 CHAPTER
FIVE
Postglacial Foragers 165 CHAPTER
SIX
The Origins of Agriculture 199 CHAPTER
SEVEN
Native North Americans 267 CHAPTER
EIGHT
Ancient Mesoamerica 321 CHAPTER
NINE
South America: The Inca and Their Predecessors 391 CHAPTER
TEN
States and Empires in Asia and Africa 437 CHAPTER
ELEVEN
Prehistoric Europe 507 iv
Brief Contents
pri31057_fm.ppi-xx.qxd
10/2/09
CHAPTER
8:03 AM
Page v
T W E LV E
The Past as Present and Future 561 APPENDIX: COMMON MEASUREMENT CONVERSIONS AND EQUIVALENTS 576 GLOSSARY
G-1
REFERENCES CREDITS INDEX
R-1
C-1 I-1
_ Brief Contents
v
pri31057_fm.ppi-xx.qxd
10/15/09
11:10 AM
Page i
Images of the Past S I X T H
E D I T I O N
T. Douglas Price University of Wisconsin–Madison
Gary M. Feinman The Field Museum
< ( : < V : y < p : W
_
pri31057_fm.ppi-xx.qxd
10/2/09
8:03 AM
Page vi
<
Contents PREFACE
xi
ABOUT THE AUTHORS
xvii
CHAPTER
TWO
The First Humans 35 CHAPTER
ONE
Principles of Archaeology 1
Introduction The Dawn of Humanity 35 CONCEPT The Family Tree SITE Hadar
Introduction 1 TIME 2 Geological Time 3 CHANGE 5 Biological Evolution 6 FUNDAMENTALS OF ARCHAEOLOGY 7 The Discovery of Archaeological Sites 8 Archaeological Excavation 11 CONTEXT, ASSOCIATION, AND PROVENIENCE 14 Analysis of Archaeological Materials 18 Interpretation of Archaeological Information 24
Images and Ideas The Basics of Archaeology 32
_ _
vi
Brief Contents
39
44
CONCEPT Dating Methods SITE Laetoli
48
50
SITE Swartkrans
52
CONCEPT Hunters or Scavengers? SITE Olduvai
55
56
CONCEPT The Leakey Family CONCEPT The First Tools
61
63
Images and Ideas Bones, Stones, and Human Behavior 66
pri31057_fm.ppi-xx.qxd
10/2/09
8:04 AM
Page vii
SITE Pincevent
138
CONCEPT Symbols and Notation SITE Lake Mungo, Australia
140
142
CONCEPT The Peopling of the Pacific
CHAPTER
THREE
CONCEPT Radiocarbon Dating
Out of Africa: Homo erectus 73
SITE Beringia
145
147
CONCEPT Pleistocene Extinction
Introduction From Hominin to Human 73
SITE Monte Verde
CONCEPT The Paleolithic Period
SITE Lindenmeier
78
151
154 157
CONCEPT Climate and Environment
SITE Kennewick Man
in the Pleistocene 81
Images and Ideas The End
SITE Zhoukoudian
of the Paleolithic 161
85
CONCEPT The First Europeans SITE Atapuerca
144
160
89
91
CONCEPT Pleistocene Mammals
95
SITE Kalambo Falls and Olorgesailie CONCEPT The Acheulean Handaxe
98 99
Images and Ideas The End
CHAPTER
of the Lower Paleolithic 101
Postglacial Foragers 165
FIVE
Introduction The World after 8000 B.C. 165 CONCEPT The Postglacial Environment
of Europe 169 SITE Vedbaek
CHAPTER
FOUR
171
CONCEPT Bone Chemistry and
Prehistoric Subsistence 176
The Hunters 105
SITE Elands Bay Cave and De Hangen
Introduction The Rise of Homo sapiens 105 CONCEPT The Origins of Language
109
SITE The Klasies River Mouth Caves CONCEPT Modern and Ancient DNA
114
SITE The Valley of the Neanderthals
117
CONCEPT The Upper Paleolithic SITE Dolni Vestonice
187
121
192
CONCEPT Contemporary Hunter-Gatherers
194
Images and Ideas The World of Hunter-Gatherers 196
123
127
SITE The Cave of Lascaux CONCEPT Portable Art
SITE Carrier Mills
183
CONCEPT The Human Skeleton
111
CONCEPT The Fate of the Neanderthals
SITE Sannai Maruyama
179
130
136
_ Contents
vii
pri31057_fm.ppi-xx.qxd
10/2/09
CHAPTER
8:04 AM
Page viii
SIX
CHAPTER
The Origins of Agriculture 199
SEVEN
Native North Americans 267
Introduction The First Farmers 199
Introduction The Diversity of Native American Life 267
CONCEPT Explaining the Origins
SITE Poverty Point
of Agriculture 205
SITE Hopewell
SITE ’Ain Mallaha
CONCEPT The Archaeology of Exchange
209
CONCEPT Wheat, Barley, Pigs, Goats,
SITE Cahokia
and Sheep 211 214 218
SITE The Draper Site
CONCEPT Archaeozoology SITE Çatalhöyük
SITE Snaketown
227
SITE Mehrgarh
at Snaketown
234
SITE Ban-po-ts’un
295
300 305
SITE Chaco Canyon
236
SITE Ozette
239 240
SITE Guilá Naquitz Cave
306
311
CONCEPT Chiefs
SITE Khok Phanom Di
CONCEPT Zea mays
293
CONCEPT Studying Community Plan
232
CONCEPT Pottery
SITE Tehuacán
224
287
289
CONCEPT Grave Offerings
221
281
283
SITE Moundville
CONCEPT Archaeobotany
CONCEPT Rice
276
CONCEPT Monumental Architecture
SITE Abu Hureyra
SITE Jericho
272
315
Images and Ideas The Clash of Worlds 317
243
246
250
SITE Guitarrero Cave
254
CONCEPT Agriculture in
Native North America 258 CONCEPT Breast-Feeding and Birth Spacing
Images and Ideas The Spread of Agriculture 262
260
CHAPTER
EIGHT
Ancient Mesoamerica 321 Introduction Early State Development in Mesoamerica 321 SITE San José Mogote
326
CONCEPT Nonresidential Architecture SITE San Lorenzo and La Venta CONCEPT The Olmec Horizon
_ _
viii
Contents
331 335
329
pri31057_fm.ppi-xx.qxd
10/2/09
SITE El Mirador
8:04 AM
Page ix
337
SITE Chavín de Huántar
CONCEPT Carved Stones and Early Writing SITE Monte Albán
342 347
SITE Sipán
CONCEPT Wetland Fields SITE Palenque
415
SITE Chan Chan
361
419
SITE Cuzco and Machu Picchu
363
CONCEPT Inca Highways
365
SITE Huánuco Pampa
CONCEPT Writing and Calendars
369
410
411
SITE Tiwanaku
357
CONCEPT Tikal’s Monument Record
SITE Tula
356
404
405
CONCEPT The Nazca Geoglyphs
349
CONCEPT The Mesoamerican Ballgame SITE Tikal
CONCEPT The Textiles of Paracas SITE Moche
CONCEPT Settlement Pattern Surveys SITE Teotihuacan
341
400
423
427
429
Images and Ideas The Organization of State Society 433
371
SITE Chichén Itzá
374
SITE Tenochtitlán
378 Pharaoh
CONCEPT Aztec Markets
383
Priests and nobility
Scribes
CONCEPT Human Sacrifice and Cannibalism
384
Artisans
Soldiers
Images and Ideas The End of Prehispanic Civilizations in Mexico 386
CHAPTER
TEN
States and Empires in Asia and Africa 437 Introduction Asia and Africa after the Transition to Agriculture 437 CHAPTER
SITE Eridu
NINE
South America: The Inca and Their Predecessors 391
441
CONCEPT Temples SITE Uruk
445
446
CONCEPT Early Writing Systems
451
Introduction Prehispanic South America 391
SITE Harappa and Mohenjo-daro
SITE El Paraíso
CONCEPT Economic Specialization
395
CONCEPT The Maritime Hypothesis
454 460
399
_ Contents
ix
pri31057_fm.ppi-xx.qxd
10/2/09
SITE Hierakonpolis
8:05 AM
Page x
462
SITE Vix
CONCEPT The Cemetery at Hierakonpolis SITE Giza and Dynastic Egypt CONCEPT Pyramids SITE An-yang
468
473
550
552
Images and Ideas Lessons from Prehistoric Europe 555
475
SITE Xianyang
CONCEPT The Bog People SITE Maiden Castle
CONCEPT The Roots of Chinese Cuisine
SITE Angkor
467
547
480
482 488
SITE Jenné-jeno
492
SITE Great Zimbabwe
496
Images and Ideas Theories of
CHAPTER
State Development 501
The Past as Present and Future 561
T W E LV E
Introduction The Past as Present and Future 561 THE VALUE OF THE PAST 561 THE HERITAGE OF THE PAST 563 CHAPTER
ELEVEN
WHO OWNS THE PAST? 564
Prehistoric Europe 507
U.S. LEGISLATION AND ARCHAEOLOGY 567 Kennewick Man 568
Introduction From the First Farmers
ETHICS IN ARCHAEOLOGY 569
to the Roman Empire 507 SITE Franchthi Cave SITE Varna
THE RESPONSIBLE ARCHAEOLOGIST 569
511
CAREERS IN ARCHAEOLOGY 572
513
CONCEPT The Iceman SITE Charavines
Images and Ideas The End 574
516
520
CONCEPT The Megaliths of Western Europe SITE Stonehenge
528
CONCEPT The Aegean Bronze Age SITE Knossos SITE Mycenae
535
_ _ Contents
G-1
REFERENCES
538
SITE Borum Eshøj
APPENDIX: Common Measurement Conversions and Equivalents 576 GLOSSARY
532
CONCEPT The Bronze Age North of the Alps
x
524
544
CREDITS 542
INDEX
C-1 I-1
R-1
pri31057_fm.ppi-xx.qxd
10/2/09
8:05 AM
Page xi
<
Preface I
mages of the Past is an introduction to prehistoric archaeology that aims to capture the excitement and visual splendor of archaeology while at the same time providing insight into current research methods, interpretations, and theories in the field. To introduce our text, we offer some background on why we wrote the book, describe its organization and distinctive features, indicate what is new in this edition, provide some information on the various supplements, and, finally, acknowledge all those individuals and institutions that have contributed to Images of the Past.
WHY WE WROTE THIS BOOK Perhaps a short history of our own teaching experience and the motivation for this volume will help convey our intent. Images is the result of a combined total of more than 50 years of teaching archaeology. Some years ago, when our introductory archaeology curriculum was revamped with an advanced section for majors and a beginning course for other students, we decided to make interesting archaeological sites the focus of a survey of world prehistory. We would discuss a series of sites from the Pliocene (now Miocene!) to the present that reflected archaeological knowledge about the past and how that information was obtained. The emphasis on sites allowed us to cover time and space and certain methods and theories, as well as provide a general survey of world prehistory. Students generally enjoyed the course and completed the semester with a broad understanding of archaeology, the major questions in the discipline, and the ways that archaeologists think about the past. But we were generally dissatisfied with the texts that were available for this introductory archaeology course. A number of introductory books on the subject already existed, of course. Such volumes generally took one of two directions: They provided either a comprehensive survey of world prehistory or a primer on method and theory. Back then, and still today, surveys of world archaeology summarized what archaeologists had learned, but they often tended to be rather dry encyclopedias of information on the many places and times that people lived in, in the past. That vast body of data is formidable to beginning students, and they have trouble
discerning what is really important. Primers on method and theory, on the other hand, were compilations of the history, techniques, concepts, and principles of archaeology: how to search for archaeological remains, how excavations are done, how to determine the age of prehistoric materials, who Louis Leakey was, and the like. Neither of these approaches met our needs, so we decided to write a text that followed the format that had been successful in our introductory class. We wrote a book that combined both survey and primer, but with an emphasis on archaeological sites. We believed then, and still do today, that a combination of what has been discovered and how archaeologists learn about the past is of value in introductory archaeology courses. We also took a new tack in the book. Rather than try to cover all of archaeology, we chose to emphasize only certain discoveries that had produced major insights into prehistory. Our focus was, and still is, on some 80 archaeological sites from a variety of times and places around the world. These sites are signposts to the past and allow students to focus on what is important. To play to our strengths, we divided up the writing according to our own areas of knowledge and activity. Doug Price is interested in prehistoric foragers and the transition to agriculture; Gary Feinman is interested in the rise of complex societies and the organization of states. Price works primarily in the Old World with stone tools, bones, and hunter-gatherers; Feinman does fieldwork largely in Mexico and China, where ceramics are a primary source of information. Our diverse interests allowed us to create a text with balanced coverage of the Eastern and Western Hemispheres. We also took a new approach to the format and layout of the text. We divided the information into pieces, with more than 80 sites and numerous small sections on ideas, methods, people, and things. These short segments, while full of information, can be readily digested by the reader and allow the instructor to organize the readings in the book as best fits the course. The substantial number of illustrations helps convey both the diversity and splendor of archaeology. Thus Images offers a visual, site-oriented look at human prehistory. What is important, we believe, is to convey the excitement, intrigue, and imagery of archaeology. _ Preface
xi
pri31057_fm.ppi-xx.qxd
10/2/09
8:05 AM
Page xii
We think that Images does that and that it provides a rich introduction to archaeology. We hope that our interest in and enthusiasm for archaeology carry over to you in this book and that you will enjoy these Images of the Past.
ORGANIZATION OF THE BOOK Our journey begins with the evidence for the first humans, more than 5 million years ago, and concludes with the rise of great empires around the world. This survey of world prehistory is organized in 12 chapters, along chronological and/or geographical lines. Chapters 2–6 are in chronological order, from the earliest human remains several million years ago to the beginnings of farming around 10,000 years ago. These chapters follow the expansion of human beings from our original home in Africa to Asia, Europe, and eventually Australia and the Americas. Chapter 6 covers the beginnings of agriculture from a global perspective. Chapters 7–11 are concerned with the rise of large, complex societies and early states and empires. This second half of the text has a geographical organization, with chapters on North America, Middle America, South America, Asia and Africa, and Europe. Within each of these chapters, we have generally followed the sequence of development through time, from earlier to later. Although the earliest state societies arose in the Old World, we have arranged the chapters from the New World to the Old in order to emphasize and compare the rise of states in both areas. This arrangement of the chapters is intended to enhance comprehension of major processes such as the spread of agriculture and the rise of more complex societies. We have sandwiched these ten chapters of siteoriented survey between two distinctive bookends that introduce the field and convey some sense of the larger context of archaeology. Chapter 1 provides a brief overview of principles and methods. This information gives the reader a basic understanding of the kinds of things that archaeologists want to know and how they find them out. Chapter 12 has been substantially expanded as a conclusion for the book. This chapter considers why archaeology is important and the ethical responsibilities of being an archaeologist.
HALLMARK FEATURES OF THE BOOK
_ _
Each of the chapters on world prehistory contains site and concept essays enclosed by an Introduction and by a concluding section called “Images and Ideas.” The Introduction provides an overview of the major themes and discoveries in the chapter. The Introduction also contains essential maps and chronological charts for the chapter. The “Images and Ideas” sections provide a recapitulation of the chapter content and place that information in a larger context, often incorporating new concepts,
xii
Preface
theories, or comparisons. Examples of discussions in the “Images and Ideas” sections are the behavioral correlates of cold climate adaptations, the origins of language, and the nature of cultural complexity. The Introductions and “Images and Ideas” sections should be read with some care; they provide the glue that binds the site descriptions together. Interspersed among the site descriptions are concept sections that cover some of the how and why of archaeology: essential methods, debates about archaeological interpretation, or certain spectacular finds. In these concept sections, we illustrate some of the more interesting questions archaeologists ask about the past and highlight various new methods that are employed to decipher the archaeological record. Because prehistory is a very visual subject, we have incorporated more than 600 illustrations in this book— more than any other book on the market. It is essential to see and study the maps, plans, artifacts, and places that help make up the archaeological record. The basic framework of archaeology is the place of prehistoric materials in time and space. For this reason, we have included a series of coordinated maps and timelines to show readers where these sites and materials fit in terms of geography and chronology. Throughout the text, we have included a number of learning aids to help students better understand the material. Chapter outlines lead off each chapter, giving students a preview of what is to come. Marginal quotes allow students to hear the voices of the field. A pronunciation guide lists difficult names and terms. Technical terms and important concepts in archaeology are indicated in bold type; these words can be found both in the adjacent margin of the text and in a glossary at the end of the book. The size and scale of archaeological sites and features is an important aspect. Where appropriate, we have tried to provide some sense of the size of areas and structures with reference to modern features such as city blocks, football fields, and buildings. An appendix offers some English–metric measure conversions and various equivalents to help make sizes more comprehensible. Supplementary readings are essential for introductory courses, for several reasons: to provide interested students with directions for further study, to assist in the preparation of papers, and to elaborate on subjects that can be addressed only briefly in a textbook. In Images of the Past, a short list of Suggested Readings appears at the end of each chapter, appropriate to the subject matter. A more complete list of sources used in the preparation of the book can be found in the back pages. Specific citations were not used in the text itself for the sake of readability, but references for the information can be found under the name of the individual associated with the work in the bibliography at the back of the book. In addition, this bibliography appears in searchable format on the book Web site.
pri31057_fm.ppi-xx.qxd
10/2/09
8:05 AM
Page xiii
An important note on dates in this edition: Because of the long time span covered by archaeology, the age of archaeological materials is given in several ways. Dates older than 10,000 years ago are described in years before the present (B.P.) or in millions of years ago (m.y.a.). Dates less than 10,000 years ago are given in calendar years before Christ (B.C.) or anno Domini, “in the year of the Lord” (A.D.). Dates for the past 10,000 years have been corrected, or calibrated, for a known error in radiocarbon dating. Another term used for more recent periods of time is millennium, a period of 1000 years. We live today in the third millennium, the third 1000-year period after Christ. The millennia before Christ run in reverse—for example, the first millennium goes from 1000 B.C. to 1 B.C., and so on.
WHAT’S NEW IN THIS EDITION This sixth edition of Images of the Past is a welcome chance to make corrections, update material, and add new material. We have retained the basic structure of a site-by-site journey through the past, interspersed with blocks of text about places, methods, and things. We believe that this connected series of short modules serves the reading habits of our students well. At the same time, we have improved the quality of the book with this round of revisions. In addition, we have substantially updated and revised all of the chapters, deleting dated material and adding new, hopefully even more interesting, information. We have made a strong effort to ensure that dates and periods are the same throughout the text. In part because of the many helpful comments and suggestions we have received, we have been able to revise this text in accordance with both new discoveries in archaeology and the interests of our readers. The past does not get old, and new discoveries and changing interpretations are a constant in archaeology. The pace of discovery and insight in modern archaeology is such that each year there are dramatic changes in our knowledge. We hope to keep Images of the Past as up-to-date as a book about the past can be. In the first part of the book, new material has been added on extinct hominids, the first Americans, genetics, the origins of agriculture, and domestication. Several sites have been deleted (Trinil, Gatecliff, and Vindolanda), as they provided less information than others. Chapter 1, on the principles of archaeology, has been modified with some new illustrations and text. More discussion of the relationship between anthropology and archaeology has been added, and more on historical archaeology as well. Context, association, and provenience in archaeology get more attention in the chapter. Phytoliths and starch grains have been added to the microbotanical remains mentioned in the chapter. And field notes are illustrated. Chapter 2, also on the first humans, has been revised slightly, with a new section on argon-argon dating,
discussion of recent discoveries of early hominins, and the latest information on Atapuerca and the first Europeans. Chapter 3 contains further mention of the socalled hobbits from the island of Flores in Indonesia. We have added material on a number of new sites, including a discussion of the discovery of fire at Gesher Benot Ya’aqov in Israel, Kennewick, Hebior, and the Jomon. In Chapter 4, several changes have been made. The sites of Pinnacle Point and Kibish in Africa have been added, in regard to the appearance of fully modern humans. Genetics continues to offer much new information— in this case, on body hair and skin color in early humans. In the Western Hemisphere, there is also news. The oldest site in North America, Paisley Cave in Oregon, contains elementary evidence of the first Americans. The Clovis phenomenon seems to have been a brief one. And a new argument that meteors caused extinctions of large game at the end of the Pleistocene is noted, but not accepted. A new box on the peopling of the Pacific has been added after the section on Australia to provide some information on this vast area of the world. Although the Pacific Islands were colonized late in human prehistory, the story of the expansion of humans across the Pacific is a remarkable one. There is a spectacular Jomon site, Sannai Maruyama, to replace the previous one. Nitrogen isotopes have been added to the discussion of paleodiet in Chapter 5. Chapter 6 contains revised art and text, with new findings on the domestication of plants and animals in Peru, the Near East, China, Africa and Europe. The term paleoethnobotany has been changed to archaeobotany in keeping with consensus in the field. It also matches well with archaeozoology and conveys the very direct connection to archaeology. The second half of the book includes new information and features. In Chapter 7, new findings concerning the introduction of corn into the American Southwest are presented. Chapter 8 presents new material on early Maya writing and new perspectives on the competition between different Maya states during the classic period. Chapter 10 incorporates new information on the mapping and size of Angkor. Chapter 11 has the latest on the Iceman, including the use of strontium isotopes to determine his place of origin. There are also exciting new data from the Stonehenge Riverside project, including interpretation of life and death contexts. Finally, we have expanded Chapter 12 again, with more material on CRM (some picked up from earlier chapters) and additional discussion of career options. We have also added new art—photos and line drawings—to improve the visual impact of Images of the Past. There are more color illustrations in the book this time, and captions for all illustrations have been expanded. Maps and timelines have been updated along with the text. We hope that these visuals convey some of the excitement of these discoveries. In addition, the Suggested Readings for each chapter have been updated. The bibliography Preface
xiii
_
pri31057_fm.ppi-xx.qxd
10/15/09
11:11 AM
Page xiv
at the end of the book has been expanded and revised. Glossary terms for each chapter have been expanded. Figures have been numbered throughout the text so that specific reference can be made to them individually. Throughout the volume, we have tried to improve the flow and accuracy of the text and to add to the connectivity of our story. We hope that the new content has resulted in a book that students will want to pick up and read. In addition to the changes in the text, the supplements for the book remain a strong component and are detailed in the next section.
SUPPLEMENTS
• Image Library—offers professors the opportunity to create custom-made, professional-looking presentations and handouts by providing electronic versions of many maps, tables, illustrations, and photos from the text. All images are ready to be used in any applicable teaching tools. • PowerPoint lecture slides—give professors readymade chapter-by-chapter presentation notes. • Instructor’s Manual—offers chapter outlines, chapter summaries, learning objectives, lecture launcher ideas, and suggested films and videos.
It also provides access to all of the student online materials. Visit our Online Learning Web site at www.mhhe .com/priceip6e to access these robust supplements.
For the Student The Student’s Online Learning Center (by Adam Wetsman, Rio Hondo College). This free, Web-based student supplement features a large number of interactive exercises and activities, helpful study tools, links, and useful information (www.mhhe.com/priceip6e). The Web site is designed specifically to complement the individual chapters of the book. In-text icons guide students to information on a particular topic that is available on the Web site. Useful Study Tools • Chapter objectives and outlines—give students signposts for understanding and recognizing key chapter content. • Multiple-choice and true/false questions—give students the opportunity to quiz themselves on chapter content and visuals. • Essay questions—allow students to explore key chapter concepts through their own writing. • Glossary—defines key terms. • Audio glossary—helps students with words that are difficult to pronounce through audio pronunciation. Useful Information • FAQs about archaeology careers—give students answers to questions on available jobs, necessary education and training, and basic texts on the field, as well as picking a college or university, going on a dig, and getting more information. • Career opportunities—offer students related links to useful information on careers in anthropology.
For the Instructor
_ _
The Instructor’s Online Learning Center (by T. Douglas Price, Gary M. Feinman, and Adam Wetsman). This indispensable, easy-to-use, password-protected instructor supplement provides a variety of features:
xiv
Preface
ACKNOWLEDGMENTS Any large project like this is the culmination of the efforts and contributions of a multitude of individuals and institutions. We want to thank the many people who have helped with this book in a number of different ways— reviewing the text, providing new data, supplying photographs and art, locating materials and information, checking facts, and giving general support. With more than 600 illustrations, the task of finding artwork, obtaining copies and permissions, and organizing it all is enormous. We have done our very best to contact the copyright holders of the original work included herein and to secure their permission to reprint their material; if we have overlooked anyone, we offer our sincere apologies. This project has been long and complex and would not have been possible or pleasurable without the help of these friends and colleagues: Kim Aaris-Sørensen, Melvin Aitkens, Niels Andersen, Larry Bartram, Gert Jan Bartstra, John Bennet, Pia Bennike, Peter Bogucki, Richard Bradley, Maggie Brandenburg, C. K. Brain, Robert Brightman, Göran Burenhult, Jim Burton, Brian Byrd, Christopher Chippendale, Tim Champion, Grahame Clark, Desmond Clark, Carmen Collazo, Meg Conkey, Lawrence Conyers, Nina Cummings, Erwin Cziesla, George Dales, Jack Davis, Hilary and Janette Deacon, John de Vos, Preben Dehlholm, Tom Dillehay, Christopher Donnan, Scott Fedick, Lisa Ferin, Kent Flannery, Melvin Fowler, George Frison, Anne Birgitte Gebauer, Henry George, Ted Gerney, Jon Gibson, Junko Habu, Peter Christian Vemming Hansen, Spencer Harrington, Sønke Hartz, Matt Hill, Ian Hodder, Brian Hoffman, Frank Hole, Vance Holliday, F. Clark Howell, Fang Hui, Tom Jacobsen, Dick Jeffries, Greg Johnson, Ken Karstens, Larry Keeley, Mark Kenoyer, Susan Kepecs, J. E. Kidder Jr., Richard Klein, François Lévèque, Katina Lillios, Henry de Lumley, Tom Lynch, Joyce Marcus, Alexander Marshack, Ray Matheny, Alan May, Roderick McIntosh, Susan McIntosh, Richard Meadow, James
pri31057_fm.ppi-xx.qxd
10/15/09
11:11 AM
Page xv
Mellaart, A. T. M. Moore, Donna J. Nash, Chris O’Brien, Inger Österholm, David Overstreet, John Parkington, Peter Vang Petersen, Tom Pleger, Theron D. Price, Naomi Pritchard, Jeffrey Quilter, John Reader, Charles Redman, Merle Greene Robertson, Gary Rollefson, Ulrick Rossing, William Ruddiman, Denise Schmandt-Besserat, Sissel Schroeder, Kathie Schick, Jeff Shokler, Brian Siegel, Ralph Solecki, Charles Spencer, Dragoslav Srejovic, Sharon Steadman, Vin Steponaitis, Jim Stoltman, J. F. Thackeray, Helmut Thieme, David Hurst Thomas, Donald Thompson, Larry Todd, B. L. Turner II, Patty Jo Watson, John Weinstein, Huang Weiwen, J. Peter White, Joyce White, Edwin Wilmsen, Peter Woodman, and Tineke van Zandt. Several individuals deserve special mention. Linda Nicholas helped greatly with many aspects of the project, especially finalizing large parts of the text and illustrations. Jennifer Blitz spent much of a year obtaining illustrations and permissions for the first edition with extraordinary energy and care. We are also very grateful to the teaching assistants we had over the years in our introductory course in archaeology at the University of Wisconsin for their input and comments. Reviewers for the sixth edition provided lots of ideas and suggestions, and we gratefully acknowledge their contributions: Charles Ewen, East Carolina University Kristy Miller, Estrella Mountain Community College Lisa Westwood, California State University–Chico Nancy Marie White, University of South Florida Slobodan Mitrovic, Brooklyn College Timothy R. Pauketat, University of Illinois Tina Thurston, SUNY Buffalo Jennifer Taschek, San Diego State University Mark A. Rees, University of Louisiana at Lafayette Lisa Frink, University of Nevada–Las Vegas Tineke Van Zandt, Pima Community College Ellen E. Bell, California State University–Stanislaus Alexia Smith, University of Connecticut Alexandre Steenhuyse, Virginia Commonwealth University Kerry Josef Pataki, Portland Community College–Sylvania We thank our reviewers for the previous five editions, who provided help, ideas, and inspiration to revise and refine the text:
J. M. Beaton, University of California–Davis Richard Blanton, Purdue University Charles A. Bollong, University of Arizona Scott Brosowske, University of Oklahoma G. A. Clark, Arizona State University Angela E. Close, University of Washington Kathryn Cruz-Uribe, Northern Arizona University Richard Effland, Mesa Community College James Enloe, University of Iowa Steven Falconer, Arizona State University Kenneth L. Feder, Central Connecticut State University Lynne Goldstein, Michigan State University William A. Haviland, University of Vermont John W. Hoopes, University of Kansas John J. Killeen, Louis Berger and Associates, Cultural Resource Group Steve Langdon, University of Alaska–Anchorage Paul E. Langwalter II, Cypress College Carole A. Mandryk, Harvard University Marilyn Masson, State University of New York–Albany Randall McGuire, State University of New York–Binghamton Alan McPheran, University of Pittsburgh Gary W. Pahl, San Francisco State University Mary Pohl, Florida State University David Pokotylo, University of British Columbia Donald A. Proulx, University of Massachusetts–Amherst John W. Rick, Stanford University Lauren W. Ritterbush, University of Kansas Ralph M. Rowlett, University of Missouri–Columbia Katharina J. Schreiber, University of California–Santa Barbara Michael P. Smyth, University of Kentucky William Turnbaugh, University of Rhode Island Peter S. Wells, University of Minnesota Adam Wetsman, Rio Hondo College Mary K. Whelan, University of Iowa Randall White, New York University
Douglas B. Bamforth, University of Colorado–Boulder
Chip Wills, University of New Mexico
Timothy Baumann, University of Missouri–St. Louis
Richard W. Yerkes, Ohio State University
David J. Wilson, Southern Methodist University
_ Preface
xv
pri31057_fm.ppi-xx.qxd
10/15/09
11:11 AM
Page xvi
This sixth edition has required a large group of talented individuals to put it together and we would like to heartily thank them. The McGraw-Hill staff included Gina Boedeker, Sponsoring Editor; Nicole Bridge, Managing Editor; David Blatty, Production Editor; Preston Thomas, Design Manager; Brian Pecko, Photo Permissions; Natalie Gibboney, Text and Art Permissions; Tandra Jorgensen, Production Supervisor; Thomas L.
_ _
xvi
Preface
Briggs, Copyeditor; Kimberly McCutcheon, Proofreader; and Janice Wiggins-Clarke, Developmental Editor. To all of these individuals go our deep and sincere thanks. We hope that you find the result worth your efforts and that you will continue to provide input, suggestions, and new discoveries that will improve the next edition. T. Douglas Price Gary M. Feinman
pri31057_fm.ppi-xx.qxd
10/2/09
8:05 AM
Page xvii
<
About the Authors
Doug Price is Weinstein Professor of European Archaeology and Director of the Laboratory for Archaeological Chemistry at the University of Wisconsin–Madison, where he has been on the faculty for more than 30 years. He is also 6th Century Chair in Archaeological Science and a part-time member of the Department of Archaeology at the University of Aberdeen. His current research involves fieldwork on the beginnings of agriculture in Denmark and lab studies using strontium isotopes in human tooth enamel to address questions of prehistoric migration. He is the author of a number of books and articles on archaeology and has been involved in fieldwork in Ireland, Wisconsin, Michigan, the Netherlands, Peru, Israel, Guatemala, Mexico, and New Mexico. He likes archaeology, most children, cooking, college football, and the family dog, Bagel. He doesn’t like long, self-promoting descriptions of the author of a book.
Gary Feinman is Curator of Mesoamerican Anthropology at The Field Museum in Chicago. He also is an Adjunct Professor of Anthropology at both the University of Illinois–Chicago and Northwestern University. Feinman’s current research, which he directs with Linda Nicholas, is focused on understanding the economy and daily life at the time of the Monte Albán state in the Valley of Oaxaca, Mexico, primarily through excavations at the site of El Palmillo. He also is involved in a regional settlement pattern project in eastern Shandong Province, China, with colleagues from The Field Museum and Shandong University. Feinman is the author of various books and articles and also has conducted field research in the North American Southwest. He has taught postgraduate classes in Mexico and China. In addition to archaeology, Feinman enjoys sports, hiking, time with family and friends, travel, and communicating about science and archaeology to the public through diverse media and means.
_ About the Authors
xvii
pri31057_fm.ppi-xx.qxd
10/2/09
8:05 AM
Page xviii
pri31057_fm.ppi-xx.qxd
10/2/09
8:05 AM
Page xix
Also Available from McGraw-Hill by T. Douglas Price Principles of Archaeology, 2007
pri31057_fm.ppi-xx.qxd
10/2/09
C
8:05 AM
H
Page xx
A
P
1
T
E
Figure 1.1 Excavations at Boxgrove, a Paleolithic site in England.
_
R
O
N
E
pri31057_ch01.pp1-33.qxd
10/1/09
7:45 AM
Page 1
Principles of Archaeology Introduction
E
xcavation is at the heart of the fascination of archaeology. Digging into the earth to reveal buried lives is an extraordinary undertaking. Excavations at the site of Boxgrove in southern England (Figure 1.1), for example, are uncovering human bones and stone tools from almost half a million years ago. Archaeology tells us about our human past. This book, Images of the Past, is about archaeology and covers more than 7 million years and much of the planet. But it is simply not possible to write about all of human prehistory in a single volume such as this; that would be like trying to see all the attractions in Washington, DC, in 10 minutes. Because we can visit only a few of the more interesting places, we have chosen important archaeological sites that have substantially increased our understanding of the past. We hope the pathway through the past that weaves through the following pages provides you with a sense of what archaeologists know about our global past and how they have come to know it. The trail that runs through this volume and ties the past to the present involves major trends in our development as a technological species—growth, diversification, and specialization. Growth is seen in the increasing number of people on the planet and in the greater complexity of human technology and organization. Diversity is observed in the variable roles and social relationships that exist in society and in the kinds of environments our species inhabits. Increasing specialization is witnessed in the tools and techniques used to obtain food and manufacture objects. The story of our human past, then, is the story of these changes over time as we evolved from small, local groups of people living close to nature to large nation-states involved in global trade, warfare, and politics. Archaeology is the study of our human past, combining the themes of time and change. Those themes—change in our biology and change in our behavior over time—are also the focus of this book. Archaeology is the closest thing we have to a time machine, taking us backward through the mists of the ages. The fog becomes thicker the farther back we go, and the windows of our time machine become more obscured. In Chapter 2, we go as far back as humans can go, some 7 million years ago, when we took our first steps in Africa. Subsequent chapters trace the achievements of our ancestors as we migrated to new continents, developed innovative technologies for coping with cold climates, crafted more complex tools, imagined art, domesticated plants and animals, moved into cities, and created written languages. But first, in this chapter, we present an introduction for comprehending our human past—those themes of time and change—along with basic methods and principles of archaeology.
1
< ( : < V : y < p : W
CHAPTER OUTLINE
Introduction 1 TIME 2 Geological Time 3 CHANGE 5 Biological Evolution 6 FUNDAMENTALS OF ARCHAEOLOGY 7 The Discovery of Archaeological Sites 8
Archaeological Excavation 11 CONTEXT, ASSOCIATION, AND PROVENIENCE 14
Analysis of Archaeological Materials 18 Interpretation of Archaeological Information 24
Images and Ideas The Basics of Archaeology 32
prehistory In general, the human
past; specifically, the time before the appearance of written records.
archaeology The study of the human past, combining the themes of time and change.
_
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 2
Figure 1.2 The Big Bang began the history of the universe, spewing space and time into the unknown.
TIME
w w w. m h h e . c o m / p r i c e i p 6 e
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
If you count one number per second, night and day, starting with 1, it would take 17 minutes to count to a thousand, 12 days to count to a million, and 32 years to count to a billion. —Carl Sagan (1987)
_ _
2
To understand time, it is necessary to imagine the unimaginable. Sometime between 10 and 15 billion years ago, an explosion of cosmic proportions ripped time and space apart and created our universe. Hydrogen and helium hurtled through the emptiness, cast out of that original Big Bang (Figure 1.2). Clouds of these gases began to coagulate, and as they were compacted by gravity, temperatures rose and the energy created in the nuclear furnaces of the first stars lit up the universe. More complex reactions in these emerging stars gave rise to heavier atoms of carbon, oxygen, magnesium, silicon, sulfur, and the other elements. Huge eruptions and disintegrations tore these early elements out of the stars and spewed them across space, creating newer and heavier stars. Smaller conglomerations of elements, lacking the mass or the temperature to ignite, condensed and gathered around the edges of the brightly burning stars. Some of these cold outliers became hard, metallic globes; others, frigid balls of gas. The planets were born. Some gases remained on the harder planets and condensed into oceans or enveloped the surface as a primordial atmosphere. Violent electrical storms, driven by energy from the stars and cataclysmic volcanic activity, rifting the surface of the forming planets, tore apart and reconstituted these elements in the early seas and atmospheres. On the planet we call Earth, formed about 4.6 billion years ago, this alchemy of primeval forces churned out new molecules in an atmosphere of methane, ammonia, hydrogen sulfide, water, and hydrogen. Among the multitude of chemistries created in the soup of the early earth’s oceans was a remarkable combination of atoms. This was a strange molecule, able to reproduce itself—to make a copy of its original—to live. Life emerged shortly after 4 billion years ago. Like the broom of the sorcerer’s apprentice in the film Fantasia, once begun, the copying process filled the seas with duplicates. These reproducing molecules grew, achieved more complex forms, and became the building blocks of more elaborate organisms that developed metabolic and sexual reproductive functions. Systems for eating and internal metabolism enabled organisms to obtain energy from other life-forms. Sexual reproduction allowed for a tremendous diversity in offspring and, thus, a greater capacity for adapting to changing environments and conditions. Plants appeared in the oceans and spread to the land. The atmosphere fed carbon dioxide to the plants, and they in turn replenished the air with oxygen through the process of photosynthesis. Swimming cooperatives of mole-
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 3
First molecules Sunday—Day 1
Bacterialike organisms
Monday—Day 2
Photosynthesis Algae
First cells Tuesday—Day 3 First cells containing genetic material Wednesday—Day 4
Invertebrates Thursday—Day 5
Trilobites Friday—Day 6 Hominins 11:48 P.M
First primates 11:00 P.M. Saturday—Day 7 Insects Swamps and forests
Dinosaurs 3:00 P.M.
Occurring within the last 5 seconds before midnight
Each day = 650 million years
First art
Agriculture and domestic animals
cules in the oceans moved onto the land and began to use the oxygen in the air for breathing and other metabolic functions. Fish, amphibians, reptiles, insects, mammals, and birds spread across the face of the earth. And then, only a moment ago in geological time, a human creature evolved as part of this great chain of living beings.
Industry
Figure 1.3 The evolution of life on earth seen as a single week in time. Planet Earth forms at 12:01 on Sunday morning, life shows up for work on Monday morning, fish evolve on Saturday morning, and the first bipedal hominins show up at 11:48 Saturday night.
Geological Time Time is a very difficult concept. The universe is perhaps 10 billion years old. Earth is roughly 4.6 billion years old. The idea of 10 billion years, 4.6 billion years, or even 1 million years is impossible for us to comprehend. But to understand our past and our place in the cosmos, we need some way to appreciate such a vast span of time. If we could compact the eons that have passed into meaningful units of time, the events of our evolutionary history might make more sense.
_ Time
3
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
era A major division of geological
time, tens or hundreds of millions of years long, usually distinguished by significant changes in the plant and animal kingdoms. Also used to denote later archaeological periods, such as the prehistoric era. m.y.a. Abbreviation for millions of years ago. epoch A subdivision of geological
_ _
time, millions of years long, representing units of eras.
4
Page 4
Consider a single week, from Sunday morning to Saturday night, as a substitute for our countdown to today (Figure 1.3). One day in this 4.6-billion-year week would represent over 650 million years, a single hour would be 25 million years, a minute would be 400,000 years, and the passage of a single second would take more than 6000 years. Roughly 4.6 billion years ago—the 7 days of our symbolic week—the earth formed in our solar system. The time was the first thing Sunday, 12:01 A.M. By Sunday evening, a primitive atmosphere and oceans had appeared, and the first molecules began to coalesce. By Monday morning, the first traces of life emerged in the shape of bacteria that evolved and multiplied. More complex bacteria and algae, using photosynthesis, began the task of converting the poisonous, primordial atmosphere to an oxygen base on Tuesday. Not until Thursday were the first cells carrying genetic material created. Late Friday morning, the first invertebrate animals—resembling jellyfish, sponges, and worms—evolved. Before dawn on Saturday morning, the seas were teeming with shell-bearing animals, such as the trilobites. Around breakfast time on Saturday, fish and small land plants appeared. By 11:00 A.M., amphibians began to move onto the land, and insects appeared in a warm landscape of swamps and forests. Late that same afternoon, the first dinosaurs crawled about. Smaller, warm-blooded dinosaurs began to produce live young and nurse them. The ancestors of modern mammals appeared shortly after 9:30 P.M. At 10:53 P.M., the common ancestor of apes and man made its home in the dense forests of Africa. The first recognizable human, walking on two legs, made an appearance at 11:48 P.M. The first art was created less than 5 seconds before midnight. Agriculture and animal domestication originated only 2 seconds before the end of the week, and the industrial revolution began just as the echoes of the last bell at midnight disappeared. To help make this huge time span comprehensible, archaeologists and geologists have developed systems for breaking the vastness of time into smaller segments. Archaeologists deal with the period of humans on the planet, roughly the past 5 or 6 million years. Archaeologists use geological time, but they also have created a means of reckoning time that reflects changes in human behavior and artifacts. This archaeological system of chronology involves divisions such as Paleolithic, Neolithic, Bronze Age, and Iron Age and is discussed in more detail in a subsequent chapter. (See Chapter 3, p. 78.) Geologists deal with the entire history of the earth and distinguish a series of eras representing major episodes, usually separated by significant changes in the plant and animal kingdoms (Figure 1.4). The Precambrian was the first major era of geological time, extending from the origin of the earth to about 600 million years ago (m.y.a.). The succeeding Paleozoic era witnessed the appearance of the first vertebrate species: fish and the first amphibians. Plants spread onto the land, and reptiles began to appear. Around 245 m.y.a., the Mesozoic era, the Age of Dinosaurs, began following a period of extinction. The Cenozoic, our current era, began about 65 m.y.a. with the expansion of modern mammals, birds, and flowering plants, following extinction of the dinosaurs. This episode of extinction is now thought to have resulted from the catastrophic impact of a meteor, causing major climatic and environmental disruption. The Cenozoic is further divided by geologists into a series of seven epochs, only the last four of which are relevant to the evolution of the human species. The Miocene, which dates from 25 to 5.5 m.y.a., witnessed the emergence of our first humanlike ancestor near the end of the epoch. The Pliocene, beginning about 5.5 m.y.a., is the geological epoch in which a variety of hominins, or humanlike creatures, appeared. The Pleistocene, beginning about 2 m.y.a., was marked by a series of major climatic fluctuations. Completely modern forms of the human species appeared toward the end of this epoch. The Recent epoch— also called the Holocene (or the Postglacial or Present Interglacial)—began only
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
Era
10/1/09
7:46 AM
Period Quaternary
Page 5
Epoch
Recent (Holocene) Pleistocene Pliocene
CENOZOIC
Miocene Tertiary
Millions of years ago (m.y.a)
Oligocene Eocene Paleocene
0.01
2.0 5.5
38 54
Early humans and giant mammals now extinct; glaciation. Anthropoid radiation and culmination of mammalian speciation. Earliest apes.
Expansion and modernization of mammals.
65 Dinosaurs dominant; marsupial and placental mammals appear; first flowering plants spread rapidly. 135
Jurassic 180 Triassic Permian
Modern genera of animals.
25
Cretaceous MESOZOIC
Important Events
245 270
Dominance of dinosaurs; first mammals and birds; insects abundant, including social forms. First dinosaurs and mammal-like reptiles, with culmination of large amphibians. Primitive reptiles replace amphibians as dominant class; glaciation. Amphibians dominant in luxuriant coal forests; first reptiles and trees.
Carboniferous 350 PALEOZOIC
Dominance of fishes; first amphibians.
Devonian 400 Silurian
440
First vertebrates, the jawless fish; invertebrates dominate the seas.
Ordovician Cambrian
Primitive fishes; invasion of land by plants and arthropods.
500
All invertebrate phyla appear and algae diversify.
540
Oldest rocks; a few multicellular invertebrates; earliest fossils at 3.6 b.y.a. Single-cell organisms appear.
PRECAMBRIAN
4600
11,000 years ago and witnessed the origins of agriculture, the first cities, and the industrial age, including our present time.
Figure 1.4 The major periods of geological time and their principal characteristics.
CHANGE Change, modification, variation—these themes describe the path of evolution from the first self-replicating molecules to fully modern humans of today. Most of the evolution of life on Earth is marked by biological evolution from one species to another in order to adapt to change. As humans, we have a second, unique system for adaptation that involves learned behaviors. Culture is a means of human adaptation based on experience, learning, and the use of tools. Cultural and biological responses to cold conditions provide an example. Humans built fires to stay warm, whereas body hair increased on other animals, such as the woolly mammoth. Within limits, culture enables us to modify and enhance our behavior without a corresponding change in our genetic makeup. As a consequence, biological evolution and natural selection alone cannot explain the culturally acquired traits of the human species. The prehistoric record of our ancestors is characterized by both biological evolution and cultural developments (Figure 1.5). Biological, rather than cultural, changes dominated our first several million years of existence. The evolution of our earliest forebears was highlighted by key changes in movement, body
culture A uniquely human means of
nonbiological adaptation; a repertoire of learned behaviors for coping with the physical and social environments.
_ Change
5
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Figure 1.5 Biological organisms and cultural artifacts change over time. The history of the automobile from A.D. 1910 to 2000. The evolution of the horse from Hyracotherium, 45 m.y.a., to Equus, 1 m.y.a.
Page 6
2000
Equus 1 m.y.a.
1970
Pliohippus 3 m.y.a. 1950
Merychippus 15 m.y.a. 1930
Mesohippus 35 m.y.a.
1910
Hyracotherium 45 m.y.a.
size, teeth, and the size and organization of the brain. The transmission of cultural traits through learning occurs much more rapidly than Darwinian evolution. The past hundred thousand years or so of our presence on the planet are marked primarily by cultural changes rather than biological ones. The story of archaeology—the search for evidence of our cultural development over time—is the subject of this book. The nature of biological evolution is briefly discussed in more detail before we return to the subject of archaeology.
Biological Evolution
y
_ _
6
What are the differences between cultural development and biological development?
The theory of natural selection, formulated by Charles Darwin and Alfred Russel Wallace in the middle of the nineteenth century, describes this process of change. Wallace and Darwin were strongly influenced by the ideas of Thomas Malthus, an English clergyman and philosopher. In his Essay on the Principle of Population (1798), Malthus observed that the growth rate of the human population potentially exceeded the amount of food available. Malthus argued that famine, war, and disease limited the size of human populations, and for those reasons the number of people did not overwhelm the resources available to feed them. In essence, Malthus noted that not everyone who was born survived to reproduce. Darwin coined the term natural selection to account for the increase in offspring of those individuals who did survive. He introduced the concept in his 1859 publication On the Origin of Species by Means of Natural Selection. During a global voyage of exploration aboard the HMS Beagle, Darwin had observed that most species of plants and animals showed a great deal of variation—that individuals were distinct and exhibited different characteristics. Following Malthus, Darwin pointed out that all organisms produce more offspring than can survive
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 7
and that the individuals that survive do so because of certain advantageous characteristics they possess. In other words, the surviving organisms are better adapted to the world that confronts them. For example, offspring with better hearing or eyesight can more effectively avoid predators. Nature’s choice of better-adapted individuals— the “survival of the fittest,” according to Darwin—leads to continual change in the species, as their more advantageous characteristics are passed genetically from one generation to the next. This basic process gave rise to the myriad creatures that occupy the world today. Evolutionary change is often described as differential reproductive success, and natural selection is the principal, though not the exclusive, mechanism responsible for it. Of course, as environmental conditions change, those physical characteristics that enhance survival and successful parenting also may vary. Views on this process of evolution change over time, too. New mechanisms for evolution have been proposed, and there is ongoing discussion about the level in populations at which selection operates, whether on groups or on individuals. There is also debate about the pace of change—whether major evolutionary modifications occurred gradually, as Darwin emphasized, or rather abruptly and suddenly. Stephen Jay Gould and Niles Eldredge of Harvard University describe the uneven pace of evolution as “punctuated equilibrium.” It now seems that some biological shifts occur gradually, as Darwin described, whereas others may occur in rapid spurts following long periods of stasis, or little change. A major theory such as evolution is modified over time, but the basic tenets of this view have withstood many tests and offer the best way to understand the emergence of life and early humans.
Does evolution happen slowly or quickly?
y
FUNDAMENTALS OF ARCHAEOLOGY As noted previously, archaeology is the study of our human past, combining the themes of time and change, using the material remains that have survived. Archaeology focuses on past human behavior and change in society over time. Archaeologists study past human culture across an enormous amount of time and space—essentially, the last several million years and all of the continents except Antarctica. In one sense, archaeology is the investigation of the choices that our ancestors made as they evolved from the first humans to the historical present. Archaeology is also a detective story, a mystery far more complex and harder to solve than most crimes. The clues to past human behavior are enigmatic—broken, decomposed, and often missing. Piecing together these bits of information to make sense of the activities of our ancestors is a challenge. This challenge—and the ingenuity, technology, and hard work necessary to solve it— creates both the excitement and the frustration of archaeology. Archaeology is a fascinating field, in part because the subject matter is highly diverse and highly human. There are so many times and places involved, and so many questions to be asked. Archaeology accommodates an extraordinarily wide range of interests: chemistry, zoology, human biology, ceramics, classics, computers, experiments, geology, history, stone tools, museums, human fossils, theory, genetics, scuba diving, and much, much more. Many of these subjects are discussed in the following chapters. Another way to regard the nature of archaeology is to consider how it fits in among academic fields of study. There are different kinds of archaeology, and disciplinary homes vary. Archaeology is usually situated in the social sciences or humanities in a university setting. In the United States, archaeology is usually part of a Department of Anthropology, which combines archaeology with biological anthropology and cultural anthropology, all focused on humans and culture. Biological anthropology is the study of the biological nature of our nearest relatives and ourselves. Biological anthropologists study bones, blood, genetics,
evolution The process of change over time resulting from shifting conditions of the physical and cultural environments, involving mechanisms of mutation and natural selection. biological anthropology The study of the biological nature of our nearest relatives and ourselves. cultural anthropology The study of
living peoples and the shared aspects of the human experience.
Fundamentals of Archaeology
_
7
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 8
growth, demography, and other aspects of living and fossil humans and primates like the monkeys and apes. Cultural, or social, anthropologists study living peoples and focus on the shared aspects of the human experience, describing both the differences and the common characteristics that exist. Archaeology in anthropology departments is sometimes designated as anthropological archaeology, or prehistory. Anthropological archaeology refers specifically to archaeological investigations that seek to answer the larger, fundamental questions about humans and human behavior that are part of anthropological enquiry. Prehistory refers to the time of humans before the written record placed us in history. Many archaeologists do study prehistory, but many also study literate societies such as the Maya and Aztec, and the urban civilizations of ancient Mesopotamia and China, where writing began. The term prehistory is often misused and applied to these early literate civilizations as well. Historical archaeology—archaeology in combination with the written record—borders on the field of history and usually refers specifically to the archaeology of civilizations of the Renaissance and industrial era.
The Discovery of Archaeological Sites
anthropological archaeology (prehistory) Archaeological investiga-
tions that seek to answer fundamental questions about humans and human behavior. historical archaeology Archaeology
in combination with the written record. artifact Any object or item created or
modified by human action. site The accumulation of artifacts
and/or ecofacts, representing a place where people lived or carried out certain activities. fieldwork The search for archaeological sites in the landscape through surveys and excavations. survey A systematic search of the landscape for artifacts and sites on the ground through aerial photography, field walking, soil analysis, and geophysical prospecting.
_ _
excavation The exposure and recording of buried materials from the past.
8
Archaeologists study change in human culture, from the time of our early ancestors to the historical present. Much of the information about the past comes from artifacts and sites. Artifacts are the objects and materials that people in the past made and used. Sites are accumulations of such artifacts, representing the places where people lived or carried out certain activities. The process of discovery, analysis, and interpretation of artifacts and sites is the basic means through which archaeologists learn about the past. Archaeological materials are most often discovered by accident. Digging and construction activities often uncover prehistoric objects; farmers and individuals in the outdoors come upon artifacts. Amateur archaeologists often know a great deal about the prehistory of their local areas and frequently find sites while walking fields. It is essential that these finds be reported to a local historical society, museum, or university. The past is too important not to share. In addition to the chance discoveries, much of the information gathering for archaeological studies requires fieldwork that is intended to locate artifacts and sites. Artifacts and sites are found either on the surface or beneath the ground. Surveys (undertaken by archaeologists to discover artifacts on the ground) and excavations (used to expose buried materials) are the primary discovery techniques of professional field archaeology. The discovery of archaeological sites depends in part on what is already known about the landscape, environment, and history of an area. Before beginning fieldwork, archaeologists check the relevant written material on the time period and place of interest. That research reveals the present state of knowledge, indicates what is not known, as well as what is, and helps establish directions for further research. Such library research is also essential to ensure that investigations similar to those planned have not already been completed. The next step is to visit the local historical society or other archaeological institutions, such as museums or university departments, where records of the area are maintained. Such institutions generally keep archives of information on the location and contents of known archaeological and historical sites. Study of those archives indicates what types of sites are already known and perhaps their size and the general content of artifacts. Conversations with local amateur archaeologists and other interested individuals can provide additional useful information. Maps are one of the most important tools for fieldwork. Topographic maps (showing the shape of the land surface with contour or elevation lines) are available for most areas and contain a great deal of information about longitude and latitude, elevation, slope, and the location of water, roads, towns, and other fea-
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 9
Figure 1.6 An aerial photograph of an effigy mound in southern Wisconsin, approximately 800 years old. The mound has been outlined in white. See also the aerial photograph of Poverty Point, Louisiana, in Figure 7.4.
tures. In the United States, the U.S. Geological Survey compiles and distributes these maps. Aerial photographs also can provide information on the location of archaeological sites (Figure 1.6). Old foundations or prehistoric agricultural fields, overgrown with vegetation and almost hidden on the surface, may appear in aerial photographs. When prehistoric structures were originally abandoned, the depressions often filled with rich topsoil, which provides better growth conditions for vegetation. In fields of wheat, for example, such different soil conditions might result in a distinctive pattern showing the outlines of houses or whole villages. In many parts of the world, such patterns are best observed from lowflying planes during a dry period in the early summer. The next step in discovering the past involves fieldwork. An archaeological survey is a systematic search of the landscape for artifacts and sites (Figure 1.7). It is not always possible to make a complete survey of the entire area under investigation, because roads, forests, other vegetation, or construction often covers substantial parts of the landscape. It may be possible to thoroughly survey only a portion of the entire area, but that portion should be representative of the larger region under investigation. The larger the proportion of the research area that can be surveyed, the better.
Figure 1.7 Field survey in Denmark. The small red flags mark the location of finds on the surface of the plowed field. This site was from the Neolithic period.
_ Fundamentals of Archaeology
9
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Figure 1.8 Working a site. Intensive surface collections are made to pick up artifacts that may help date the site. One archaeologist holds a stadia rod used to measure the elevation.
_ _
10
Page 10
The basic type of archaeological field survey involves systematic field walking. Field crews walk up and down cultivated fields and exposed surfaces. The intervals between the walks are determined by the size of the sites that may be in the area and the nature of the ground cover. When an artifact is found, it is put in a bag, and the location of the find is recorded (Figure 1.8). The surrounding area should be searched carefully by walking back and forth at close intervals. It is important to determine whether the object is a single, isolated find or whether there are more artifacts. Surveyors also look for unusual discolorations on the surface that might indicate features such as fireplaces or pits. It is important to establish the area covered by artifacts to determine the size of the site and to obtain an estimate of the density of artifacts. Information must be recorded about each find. These field notes should include such information as (1) location, site number, map number, which field, and position in the field; (2) the archaeological material found: types and number of artifacts, fire-cracked stones, charcoal, and so on; and (3) observations about the site—for example, discolorations in the soil that could indicate cultural layers or pits, the presence of mounds, stone foundations or walls, nearby streams or other sources of water, and other pertinent environmental information. Archaeological remains are often buried beneath the sediments that have accumulated since their deposition. Objects found on the surface often have been brought up from deeper layers through agricultural or animal activities. Such materials usually provide only a partial indication of the information that can be obtained from a buried site. Once buried sites have been located by survey and have been mapped, other kinds of fieldwork can be undertaken to learn more about them. Boring into the ground with an auger or corer brings up a column of soil showing the sequence of layers and samples of sediments at the site. Small test pits, perhaps 1 × 1 m in size, dug into the ground can provide similar information and may be necessary to determine if a buried site is present. A number of borings and/or test pits often are made, following a regular pattern over the surface of the site. Soil samples should be collected from all parts of the site and at varying depths. Physical and chemical analysis of soil samples may provide information about the origins of the deposits, the water content and fertility of the soil, the amount of organic material, and the basic chemistry of the soil. These studies may provide further information on environmental and human activities involved in the formation and burial of the site and help to explain the conditions of preservation. Phosphate analysis of the sediments from a site may reveal traces of human activities. Phosphate is found in bone, feces, urine, and other organic matters that accumulate in and around human habitation. Phosphate appears as a strong blue color in the soil sample when hydrochloric acid and ascorbic acid are added. Areas with higher concentrations of phosphate show up as stronger blue colors in such analyses. Phosphate testing may supplement surface surveys in areas where vegetation prevents observations of the surface or where cultural layers are buried deep under the topsoil. Within a known habitation area, these tests may be used to determine the extent of the site and to detect special areas such as house floors. Other objects in the soil also are informative. Materials found in soil samples often include pieces of wood and plants, seeds, fragments of insects, mollusk shells, hair, or chips of bone or stone. Such items provide information on the formation of the layers, the local environment, and the nature of past human activities. For example, if small chips that result from the manufacture of arrowheads and other stone tools are present in borings and test pits, it is likely that tools were made or used in the vicinity and that other buried artifacts are present. Geophysical prospecting can be used to detect disturbances in the subsoil and the presence of prehistoric features. These methods include measurements of magnetic variations in the ground and of the electrical conductivity (resistivity) of the soil, and the use of ground-penetrating radar. Metal detectors, for example,
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 11
Antenna
Wall (a)
(b)
register the presence of metal objects on the surface and buried in the soil. Metal detectors emit an electromagnetic field that is disrupted by the presence of metal objects in the ground. Magnetometers can provide a map of the magnetic anomalies in the ground and are very useful for finding buried structures. The use of ground-penetrating radar (GPR or georadar) is standard practice on many archaeological excavations to look for features and structures before excavation. The use of ground-penetrating radar is a technique for studying buried archaeological sites (Figure 1.9). Electromagnetic waves in the form of georadar are sent into the ground, something like the sonar used in submarine hunts. Lowenergy radar waves register anomalies in the subsoil, which are shown on a map or a graph. Excavation is often required to identify such irregularities. In sum, prehistoric sites are often found through a combination of archival research and fieldwork. Archival research provides information on what is already known about an area. Fieldwork often results in the discovery of the unknown. When new sites are discovered, surface survey, testing, boring, and several geophysical methods are available to determine the size and possible contents of the prehistoric deposit. However, once a site is discovered and defined from the surface, excavations are often necessary to expose what lies beneath the surface of the ground.
Archaeological Excavation Excavation is the technique that archaeologists use to uncover buried remains from the past. Buried materials usually are more abundant and better preserved than those found on the surface. In excavations, accurate information can be observed on the arrangement and relationships of structures, artifacts, plant and animal remains, and other materials. The term in situ (Latin, “in place”) is used to describe archaeological remains in their original position of deposition. Excavation often is essential to obtain more information about the past. Excavations are conducted to answer specific questions that the archaeologist would like to answer: Who lived at the site? What did they eat? What did they do? Where did they get raw materials for making tools and equipment? What kinds of relations did they have with their neighbors? How was their society organized and structured? How did they understand the world around them? and so on. The Excavation Director The direction of an excavation requires a variety of skills and knowledge for planning the field season, raising money to pay for the work, supervising and training a crew of volunteers or students, recording the
(c)
Figure 1.9 Georadar in action. (a) Lawrence Conyers and assistant pulling the ground-penetrating radar (GPR) across an open area at Petra. (b) Schematic drawing of the instrument in use, emitting microwaves and measuring the response with an antenna. (c) A computergenerated display of the results of the magnetometer survey showing the outline of a rectangular structure buried in the middle of the open area. Test pits at this location revealed that stone walls were being recorded by the GPR.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the opening of the tomb of Tutankhamen, see the Internet exercises on your online learning center.
ground-penetrating radar (GPR or georadar) An instrument for re-
mote sensing or prospecting for buried structures using radar maps of subsoil features.
Fundamentals of Archaeology
11
_
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 12
information from the site with drawings and photographs, and measuring and mapping the location of all finds, samples, and features (Figure 1.10). The director must monitor progress in the field laboratory as well, where finds are washed, sorted, cataloged, and bagged for storage. Some knowledge of preservation techniques is necessary to conserve fragile objects. Excavations require reams of drawings, recordings, and other paperwork. The director must keep an excavation log or diary, recording the course of the excavations, the work schedule, the number of people working, accounts of expenses, dimensions and positioning of excavation areas, layout of the measuring system, and all finds. There must be recording systems for all measurements, for observations and interpretations, and for all drawings, photos, and samples.
Figure 1.10 Excavations at a Mesolithic site in Denmark. Measuring, recording, studying.
The Field Crew Archaeology is the science of the past, but it is also a social experience in the present. Excavation is a labor-intensive undertaking, and the field crew is the most important part of the project. This crew is a group of individuals involved in the actual digging process, unearthing the sites and artifacts. Crews are composed of a variety of individuals, young and old, ranging from professional archaeologists with advanced degrees to undergraduate and graduate students, and sometimes people just interested in the subject. Fieldwork can require a few days, weeks, or months and can involve walking miles each day with one’s head down in a survey of the ground or moving tons of earth to expose buried levels. Excavations are hard work, often in the hot sun. Frequently, they are carried out in remote places, requiring patience and endurance. Archaeology is also good dirty fun, and the experience of working, and relaxing, with others who enjoy the same things can be unforgettable. The discovery process is captivating, and sharing that excitement with colleagues and comrades enhances the entire experience. Fieldwork is, finally, an extraordinary learning experience. One realizes the difficulties involved in recovering information from the past and comes to appreciate what has been previously learned. In addition, a constant stream of questions about the past and the significance of place, artifact, and context comes to mind during the process. All in all, archaeological fieldwork can be one of the most stimulating activities there is. Selecting Sites for Excavation The choice of which site to excavate is determined by several factors, including potential danger to the archaeological remains. Archaeological sites are being destroyed at a rapid rate by the growth and development of modern civilization, and there is a serious and real concern about the loss of undisturbed sites for future research. Sites threatened by
Figure 1.11 Archaeological field notes: two pages of information on an Alaskan koniag house and its features.
_ _
12
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 13
THEN Figure 1.12 The vagaries of preservation. Organic materials—wood, bones, features, antlers, hides, and the like—rarely survive in archaeological sites. The upper drawing shows some original material from the Stone Age in Scandinavia, including fishing and hunting spears, fishing nets, clubs and axes, a bow and arrows, baskets and bags, necklaces and pendants, and other tools. The lower drawing shows what would remain after the organic material decayed: the stone arrowheads, an axe, and the stone weights for the fishing net—only a tiny part of the total equipment that was in use in the past.
NOW
modern construction are often good candidates for excavation. The rescue excavation, intended to save information from such sites, is the most common type of project taking place today. Sites are also chosen for excavation because they appear to be well preserved or to contain new information that will help us to better understand the prehistory of a particular region. The choice of a site for excavation is often based on the results of a survey. An initial survey of an area, including coring and testing, may indicate that one or several sites would be worth excavating. Careful surface collection and testing must be carried out at the site selected for excavation to make sure the site can provide the kinds of information that are needed and to assist in planning the excavation. Historical archives may be studied over and over again, but archaeological sites are nonrenewable resources, something like endangered species. Excavations involve moving the earth and all its contents from a site. Every excavation means the destruction of all or part of an archaeological site. All that is left when an excavation is over are the finds themselves, the unexcavated parts of the site, and the samples, photographs, drawings, measurements, and other notes that the archaeologists made. Accurate notes and records of the layers, structures, and artifacts at a site are essential, not only for the investigator, but also to create a permanent archive of information about the site that is available to others (Figure 1.11).
_ Fundamentals of Archaeology
13
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 14
The contents of a particular site are a matter of preservation (Figure 1.12). Important factors in preservation include the age of the site, the effects of erosion and deposition, bioturbation, and conditions of humidity and acidity. Archaeological sites vary from excellent conditions of almost complete preservation in extremely wet or arid conditions to poor acidic situations where almost nothing is left but inorganic objects of stone, pottery, and perhaps charcoal. Examples of extraordinary preservation can be seen in the Tollund Man from the waterlogged bogs of Denmark (see Figure 11.49) and the Iceman from the frozen glaciers of Alpine Italy (see Figure 11.10). Very old sites from the Paleolithic rarely have good conditions for preservation, and thus only stone tools and occasionally bones are preserved. It is important to know as much as possible about a site before full-scale excavation in order to choose the best strategy for the project. At every excavation, the archaeologist is faced with a series of decisions about how to achieve the most and best-documented information. Under ideal circumstances, a site could be fully excavated and everything recorded in the finest detail. In the real world, however, constraints on time and funding and a need to leave a portion of the site for future archaeologists make it standard practice to excavate only a part of the total site.
Figure 1.13 A total station in use mapping an archaeological site in the highlands of Peru. Two members of the field crew work at the total station, and two others are locating and marking map points with the reflecting target for the total station. The total station uses a laser beam to measure the distance and angle between the instrument and the target and then calculates the exact position of the target. w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the innovative area of GIS, see the Internet exercises on your online learning center.
bioturbation Activities of plants and animals in the earth, causing disturbance of archaeological materials. total station A computerized survey-
ing and mapping instrument that uses a laser beam or radio waves to measure the distance and angle between the instrument and the target and then calculates the exact position of the target. archaeological record The body of
material and information that survives for archaeologists to study.
_ _
context The association and relationships between archaeological objects that are in the same place.
14
Maps and Grids Accurate mapping of layers and artifacts is the key to the proper recording of information at an archaeological excavation. The exact topography, or shape, of the site must be recorded in the form of an accurate contour map made using a surveyor’s level and the site grid. A grid is marked out across the surface of a site before excavation. This grid should be used for all horizontal measurements. A site grid represents a coordinate system, usually with lines running north-south and east-west at regular intervals. Intervals along the two axes of the grid are designated with a system of letters or numbers or both. The grid lines and measurements within each grid square are measured as distances in meters and centimeters north and east of the baselines at the edge of the excavations. The site grid may also be oriented according to local topography or archaeological features such as mounds or middens. At coastal sites, trenches are sometimes excavated perpendicular to the coastline to study layers and site formation in relation to the coast. In a narrow cave, the grid is often aligned to the long axis of the cave. Location of the site and the site grid in relation to global latitude and longitude must be determined. A control point, or site datum, must be located in the neighborhood of the excavation as a point of origin for vertical measurements. A preexisting datum point, such as a surveyor’s benchmark, may be used if available. Otherwise, a permanent feature, such as a rock outcrop or a building foundation, may be marked and used as the datum point. The location and elevation of this point must be established in relation to known points, such as geographic features or distant benchmarks. Vertical location in the excavation is best determined using a surveying instrument, set at a known elevation, and sighting on a vertical measuring rod. Measurements at the site should be converted to meters above sea level, or the elevation of the datum line may simply be recorded. In archaeology today, a total station is normally used to electronically map the site, record elevations, and determine the location of architecture, features, and artifacts (Figure 1.13).
CONTEXT, ASSOCIATION, AND PROVENIENCE The body of evidence that archaeologists work with is part of the archaeological record—the information about the past that has survived to the present. This record includes both past materials and the context in which they are found. Context is an essential aspect of archaeological information. Context involves the association and relationships between objects that are in the same place.
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 15
At a basic level, context concerns relationships among artifacts. Items that are found together in the same pit, the same layer, or the same sediment, for example, are assumed to be related in terms of time and activity. That is, objects in the same context are thought to have been in use together in time and geographic space, roughly contemporary, and involved in the same activities or resulting from similar behaviors. In a broader sense, context is the physical setting, location, and association of artifacts and features. Context is of major importance in archaeology and provides much of the information necessary for the determination of authenticity and significance. Context is essential for learning about age, use, and meaning. The more that is known about the context of archaeological remains, the more that can be learned about the past, of both the artifacts themselves and the people who made them. A distinction is made between primary and secondary context. An object in its original position of discard or deposition, in the place where it was left, is said to be in primary context or in situ (Latin, “in place”). Objects that have been moved from their original place of deposition are in secondary context and so are less useful for learning about the past. When artifacts are removed from their original location, without proper excavation and documentation, contextual information is lost forever. Looters are unconcerned with the context of archaeological materials. Peter Cannon-Brookes describes looted artifacts as “cultural orphans, which, torn from their contexts, remain forever dumb and virtually useless for scholarly purposes.” An important term in the realm of context is provenience, or place of origin. The provenience of an artifact—the place it was found—is very important. Provenience implies context, meaning that there is additional information available about the object of interest. Artifacts and other archaeological objects with an unknown provenience provide very little information for learning about the past. Test Pits Preliminary examination of a site involves digging a series of one or more trenches or small, vertical test pits, perhaps 1 × 1 m in size, across the site. The test squares to be excavated may also be placed in rows or in a chessboardlike pattern across the site. Alternatively, their location may be chosen at random. The size and the number of test pits to be excavated depend on the kind of information being sought. In some cases, it is difficult to visualize the stratigraphy, or set of layers, observed in the small test pits. One or two long trenches across the site may provide a better view of the stratigraphy. Vertical Excavations Excavations are generally either vertical or horizontal. Vertical excavation takes the form of test pits or trenches carefully placed across a site to expose the stratigraphy and artifact contents of a site (Figure 1.14). By studying the vertical walls (the sections) of such pits or trenches, archaeologists can identify stratified layers of soil sediments. The stratigraphy, or layers, of natural sediments and human deposits reveals how the site was formed and how materials accumulated (Figure 1.15). The relationships between deposits in the stratigraphic sequence indicate the chronological arrangement of the layers. The bottom layer is deposited first as the oldest layer in the sequence. The subsequent layers are progressively younger—the law of superposition. The stratigraphic sequence provides a relative chronology whereby layers and the artifacts they contain can be determined to be “younger than” or “older than” other layers and artifacts in the same sequence. The thickness of a layer is determined not so much by the length of time that it took to accumulate as by the natural and human activities involved in the deposition of the materials. Heaps of shells may accumulate very rapidly into high shell middens (large dumps of shells from mussels, oysters, or other species); the collapse of houses with earth or sod walls results in very thick layers; stone
Figure 1.14 A section of an excavation trench exposing a stratigraphy of stream and lake deposits that succeeded one another as water levels changed in this area. The upper part of the deposit is recent blown sand.
What are some of the important skills an archaeologist needs?
y
primary context (in situ) An object found where it was originally located in antiquity, not redeposited. provenience The place of origin for archaeological materials, including location, association, and context. shell midden A mound of shells ac-
cumulated from human collection, consumption, and disposal; a dump of shells from oysters, clams, mussels, or other species found along coasts and rivers, usually dating to the Holocene.
Context, Association, and Provenience
15
_
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Figure 1.15 Vertical excavation. Archaeological sites often were places of repeated human occupation. In this artist’s reconstruction, several periods of settlement at the site of Coppergate in York, England, are shown in their stratigraphic context. The use of the site goes back almost 2000 years.
Page 16
Coppergate Site 1. Modern (1750–present) 2. Post-Medieval (1550–1750) 3. Medieval (1200–1550) 4. Norman (1067–1200) 5. Viking III (975–1067) 6. Viking II (925–975) 7. Viking I (850–925) 8. Anglian (400–850) 9. Roman (A.D. 71–400)
2
1
3 4 5 6 7 8 9
tool manufacture can produce extensive debris. On the other hand, the place where an animal was killed and butchered—a kill site—may leave almost no archaeological trace. Evaluation of a stratigraphic sequence involves distinguishing between natural and human activities. Such environmental factors as soil erosion or flood deposits may add to the local accumulation but may also remove part of a layer. Younger features such as postholes or storage pits may have been dug into older deposits. Relationships between layers must be studied carefully to determine whether younger deposits are cut into older layers and whether animal activities, downed trees, floodwaters, or later construction has disturbed or destroyed the original stratigraphy. Assessment of the context and relative position of layers allows an archaeologist to interpret the depositional history from the stratigraphic sequence. An actual calendar date of the layers may be derived from artifacts with a
_ _
16
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 17
known date found in a particular layer. For example, a hubcap from a 1935 Ford would indicate that the layer could not have been formed before 1935. The ages of many types of pottery and stone tools are known and can be used to suggest an approximate date for archaeological levels. Layers and artifacts also may be dated by means of such absolute techniques as radiocarbon and other dating methods (see p. 145).
It takes very special qualities to devote one’s life to problems with no attainable solutions and to poking around in dead people’s garbage: Words like “masochistic,” “nosy,” and “completely batty” spring to mind. —Paul Bahn (1989)
Horizontal, or Area, Excavations Horizontal, or area, excavations are often the next step after initial vertical trenches reveal structures or features to be uncovered. Such excavations expose large areas of ground, one layer at a time. These horizontal layers are recorded and removed individually. Area excavations are intended to recover information on site arrangement and structures. Such excavations may expose actual prehistoric living floors and structures where a group of people carried out everyday activities (Figure 1.16). When the site stratigraphy is relatively simple—with only one or two stages of occupation and thin cultural layers—it is possible to separate the remains from each stage of occupation. In such cases, it is advantageous to expose large surfaces of the same layer to get an overview of the distribution of features and artifacts at the settlement. Following removal of the topsoil, the surface is scraped with trowels or shovels, loose soil is removed, and features and artifacts are uncovered. The uncovered surface is then carefully recorded, usually in drawings and photographs. The sediments removed during the excavation are normally shaken or washed through fine screens to recover smaller items such as bone fragments and plant remains that otherwise may be missed (Figure 1.17). Various kinds of samples are taken from different layers in the walls of the sections and from the occupation floor. The excavated soils are usually sifted through screens and/or washed with water to find even the smallest objects, fragments of bone, and plant remains. Soil samples are taken to help define and characterize the deposits at the site. Pollen samples are sometimes collected to assist in defining the vegetation in and around the site. At most sites, samples of charcoal and bone are taken for radiocarbon dating. After removal of one layer of soil and artifacts, the procedure is repeated and a new surface is uncovered and recorded. One strategy for maintaining control of the stratigraphy is to leave a number of narrow sections untouched in the
Figure 1.16 Horizontal excavation at a prehistoric village in Wisconsin. Only one-half of a feature is excavated at first. This feature, a house depression, is being excavated with trowels. The small wall in the middle is kept as a record of the feature’s stratigraphy.
_ Context, Association, and Provenience
17
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 18
Figure 1.17 Soil from the excavation is sifted through a screen to recover small objects.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on dating of archaeological material, see the Internet exercises on your online learning center.
excavation area. The walls of these sections are cleaned and studied as the excavation goes deeper. This kind of excavation aims at recording and then removing each horizontal layer individually. Toward the end of the excavation, the sections are excavated. The surface of the sediments beneath the occupation layer is uncovered and cleaned. Unusual colors in the soil may reveal features such as pits and postholes, which are recorded by photos and drawings. Features are dissected by excavating one-quarter or one-half of the pit or posthole at a time to remove the contents and determine the function of the feature (Figure 1.18). This produces a vertical section through the middle of the feature. Sections of all features are recorded by photos, drawings, and soil description. Postholes belonging to the same structure are grouped by examining the depth of the holes and the kind of soil present. Other features are studied to determine their function and mutual relationship. At the end of the dig, the excavated area has to be filled up and undisturbed portions of the site protected in the best possible manner.
Analysis of Archaeological Materials Analysis of recovered artifacts may begin concurrently with the fieldwork in a field laboratory, or records, artifacts, and samples may be shipped back to a home laboratory to be cleaned, cataloged, and prepared for analysis. More fragile objects will require careful conservation to protect them and ensure that they do not disintegrate. After the fieldwork come more detailed analyses of the recovered materials, the writing of excavation reports, and the preparation of publications, all of which require much more work and time than the excavation itself. One estimate suggests 5 weeks of analysis and writing for each week spent in excavation. Final results of the investigations are made available to the public and to professional archaeologists through articles in scientific journals, in published reports, and in books. Archaeological fieldwork produces several major categories of finds and information: (1) artifacts—portable objects altered by human activity; (2) ecofacts— the remains of plants, animals, sediments, and other unmodified materials that
_ _
18
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 19
Figure 1.18 Excavations expose postholes and the foundations of two houses, one built over the other. The dark rectangles mark the location of construction posts. Two fireplace pits can be seen at the center and right of the photo.
result from human activity; (3) features—the immovable structures, layers, pits, and posts in the ground; and (4) sites and settlements—the set of artifacts, ecofacts, and features that defines places in the landscape where activity and residence were focused. A variety of specialists are needed in archaeology to examine and interpret the wide range of materials and information that is found at archaeological sites. There are specialists in archaeobotany, archaeometry, archaeozoology, bioarchaeology, classical archaeology, geoarchaeology, historical archaeology, paleoanthropology, theoretical archaeology, underwater archaeology, and many others. There are also specialists in certain classes of materials. For example, lithic specialists analyze the stone tools that are often a common object at archaeological sites. Ceramic specialists study the sherds of ancient pottery. Archaeobotanists (also known as paleoethnobotanists) study the plant remains, both visible and microscopic, that are found at a site. Archaeozoologists investigate the animal bones that represent the remains of meals and manufacturing activities. Bioarchaeologists are often trained in both archaeology and biological anthropology; they describe and interpret the human bones and teeth that may be found. Paleoanthropologists are archaeologists and physical anthropologists focused on the very earliest human fossils and artifacts. Geoarchaeologists and micromorphologists investigate the geological setting of sites and the details of the sediments encasing archaeological remains. Archaeometrists date those remains and undertake the chemical characterization of prehistoric materials to learn about their composition and source. Historical archaeologists and classical archaeologists work in specific time periods, with historical documents and with the classical civilizations of the Mediterranean (Rome, Greece, and others). Underwater archaeologists focus on shipwrecks and submerged archaeological sites (Figure 1.19). Artifacts Each object from the excavations must be washed to remove dust and dirt (Figure 1.20). At some sites, each object is recorded by number in a catalog. At other sites, artifacts are recorded by material and context or by the excavation area where they were found. Numbering artifacts with permanent ink ensures that each item has a label with information on the site and location of the find (Figure 1.21). The catalog description of each artifact includes a record of the kind of artifact, the type of raw material, the color, the overall shape and measurements, techniques of manufacturing, presumed function, decoration, and provenience
Figure 1.19 Underwater archaeology—a growing part of fieldwork. Divers discover and excavate a variety of finds beneath the sea, including individual artifacts, shipwrecks, and entire settlements. In this case, divers in northern Germany are at work excavating a Mesolithic settlement from approximately 6000 years before the present.
ecofact Any of the remains of plants,
animals, sediments, or other unmodified materials that result from human activity. feature An immovable structure or layer, pit, or post in the ground having archaeological significance.
Context, Association, and Provenience
19
_
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 20
information. This description could be supplemented with an accurate drawing and a photograph of the artifact (Figure 1.22). An inventory of the materials from the excavation then can be made by counting and recording the number of artifacts in each category of material, such as chipped stone, ground stone, or pottery. Following this initial recording, artifacts are classified into other categories and types. Classification is a way of creating order in a mass of archaeological materials by dividing objects into groups on the basis of shared characteristics. One example of such classification is the initial division of the remains into artifacts, ecofacts, and features, described earlier. Another example is the division of chipped stone artifacts into axes, scrapers, knives, and arrowheads. Three primary attributes are used to classify archaeological artifacts: (1) form— the size and basic shape of the object; (2) technology—the characteristics of raw material and manufacturing technique; and (3) style—the color, texture, and decoration of the object. Most of this information is recorded in the laboratory after the artifacts have been cleaned and cataloged.
Figure 1.20 Items from the excavation are washed, dried, and put in bags with labels showing their location in the site.
Ecofacts Ecofacts are unmodified natural items, such as animal bones and plant remains, that are usually brought to the site by its occupants and useful for the study of past human activity. They are used to reconstruct the environment of the site and the range of resources that people used. Ecofacts are classified as organic (plants and animals) or inorganic (sediments and stone). These materials are usually studied by archaeologists or specialists with training in botany, zoology, or geology. Plant remains from an archaeological site may include pollen, seeds, leaves, pieces of wood, and the like, depending on the quality of preservation. Visible (macroscopic) and invisible (microscopic) plant remains are distinguished. Invisible remains are often more likely to be preserved. Microscopic remains include pollen, phytoliths, starch grains, and other materials. Each type of plant produces distinctively different-looking pollen. Because of its long-distance distribution, pollen is likely to reflect the total environment around the site. Phytoliths are tiny pieces of the plant “skeleton” composed of silica. These plant parts can survive thousands of years and are often recognizable as to their
Figure 1.21 Sorting, numbering, and cataloging artifacts and other finds from an excavation are often a long and complicated process.
_ _
20
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 21
species. Many kinds of plants produce distinctive starch grains that can be preserved to the present. Changes in the types of pollen at a site over time can be used to reconstruct the vegetation history of the area and to provide a record of climatic changes. Special growing requirements and other characteristics of certain plants may reflect certain climatic conditions or specific local situations such as an open rather than a forested environment around a site. Macrofossils of botanical materials are visible remains, such as seeds and plant parts, that are more likely to be present at a site because of direct human utilization. Identification of these remains indicates what species of plants were present, whether they were wild or domestic, and in what context they were found. It is important to study the context of these remains to know how the plants were used. Plants may be collected for food, but they may also be used for production of textiles, mats, and baskets; for the manufacture of poison for arrowheads; or as drugs. Types of plants and their growing condition may also provide an indicator of the nature of the local environment and climate. Faunal analysis, or archaeozoology, is the term used to describe studies of the animal remains from archaeological sites. Animal remains are tabulated by the kinds of bones, teeth, antler, and horn that are present (Figure 1.23). Species are identified and the numbers of individuals of each species are calculated. These studies show what animals were hunted and eaten and in what proportion. The amount of meat available from each animal also may be calculated to determine the animals’ relative importance in the diet. Faunal analysis also can provide an estimate of the ratio of adult to juvenile animals and of male to female animals. A predominance of certain age groups in a species such as deer may indicate that seasonal or selective hunting was practiced. For example, a site that contained a large proportion of 3- to 6-month-old deer would suggest the animals were killed primarily in the fall, since deer are born in the spring. The presence or absence of certain parts of the animal skeleton may indicate the way animals were butchered and whether they were dismembered on the spot or killed elsewhere and selected steaks and chops brought back to the settlement. Not all animals are necessarily hunted for food. Nonfood items such as antler, fur, bone, and hides also are important materials from hunted animals.
Figure 1.22 Recording the shape of flaked stone artifacts.
Figure 1.23 Bones and teeth from an archaeological excavation, including domesticated sheep and pig.
_ Context, Association, and Provenience
21
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 22
Many different kinds of tools and equipment are made from animal products (Figure 1.24). Bone was a very important material for prehistoric peoples. The most important inorganic ecofacts are the various sediments uncovered by excavation. Deposits of soils and sediments at human settlements result from both human and natural processes. These sediments and deposits are studied by geoarchaeologists. The types of sediments present may indicate the source of the material that was deposited. Examples include water-lain silts from a flood, volcanic ashes, and frost-cracked rocks from the ceiling of a cave. The study of soil chemistry is an important aspect of the analysis of soils and sediments.
Figure 1.24 A bone point, Mesolithic Europe, 5000 B.C. This example is made from the leg bone of a small deer, split, ground, and polished to a point. It was then lashed to a wooden shaft for spearing fish.
_ _
Features Features must be studied largely in the field, since they are fixed in the ground. Features may be structures such as houses or pits, or fences or field systems defining an area used for special purposes, or constructions for certain activities, such as drying racks, fireplaces, and traps. They are useful for understanding the distribution and organization of human activities at a site. For example, the size, elaboration, and location of dwellings or burials may suggest differences in wealth and status. Some features result from the accumulation of garbage and debris, rather than from intentional construction. They include shell middens, heaps of waste material in workshops, and quarries. Studies of these features may indicate strategies for obtaining food or raw material, how the raw material was used and distributed, and whether it was scarce or abundant.
Figure 1.25 Two Viking-age burials from Denmark, A.D. 1000. These individuals were beheaded and buried with a bad view.
22
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 23
Burials and human bones are a special category of feature often found at archaeological sites. Several kinds of burials can be found (Figure 1.25). Simple inhumations represent the laid-out burial of the whole body. Such graves usually contain an articulated skeleton with all the bones in their correct anatomical positions. Secondary burials are the result of burial of some of the skeleton, after the flesh and soft tissue have disappeared. Usually, the skull and the larger bones are present, often in a small pile or bundle. Cremations are burials of the ash and small, carbonized bones from bodies that have been burned. Bioarchaeologists identify and analyze human remains. The sex of the skeleton can be determined by examining the size and shape of the pelvis and the skull and the thickness of the bones. The age at death can be estimated by the eruption sequence and wear of the teeth, the fusion (closing) of sutures between bones of the skull, and the fusion of the ends of the limb bones to the shaft. The health status of past populations can be investigated by recording the incidence of trauma that affects the skeleton. Such diseases and injuries include bone fractures, arthritis, and periodontal diseases. Nutritional problems may be reflected in poorly developed bones and the low average height of the population. Cultural practices such as cranial deformation and dental mutilation (practiced in prehistoric America, for example) also show up in the skeletal remains. Sites and Settlements Settlement archaeology is the study of how and why prehistoric remains are distributed across the landscape. Investigations range from the analysis of the location of different activities within a single room to the distribution of sites in a region. There are at least three levels of locational information: (1) a room, a structure, or some other occupation surface, such as a cave floor; (2) a site or settlement; and (3) a series of sites within a larger region. The spatial organization within a single structure defines areas for special activities such as grinding flour, cooking, weaving, or manufacturing tools, or for certain facilities such as those for sleeping or storage. Study of such organization may indicate a division of male and female space and activities, the number of people in a household, and the structure of the family—nuclear, extended, or polygynous, for example. A settlement generally includes a habitation area with one or more houses and fireplaces; different activity areas for food preparation, curing of animal skins and hides, the manufacture of various artifacts, and perhaps storage equipment; and a midden or trash area. Spatial patterning within a site can provide information about the number of houses and people at the settlement and about their relationships with one another. In addition, most of the day-to-day activities of the occupants should be reflected in the various structures and activity areas found throughout the settlement (Figure 1.26). Structures at a site may be solid and substantial in the case of permanently settled communities in a village or townlike setting. Short-term or seasonal settlements, however, may leave little trace of construction. The size of a settlement in horizontal and vertical extent depends on the number of people who lived there, the length of time they lived there, and the kinds of activities that took place and structures that were erected, as well as environmental factors. Sites of similar size could have been created by a few permanently settled people or through the occasional use of the same spot by a larger group of people. Differences in the size and architectural elaboration of houses may be evidence of status differentiation, a situation in which some people have more wealth and control over goods and labor than others. The arrangement of houses at a settlement also may reflect social organization in the separation of poor and wealthy households. Concerns for privacy and protection in the form of fences, palisades, or ditches may indicate private ownership or conditions of competition or warfare. In addition, settlement studies may reveal areas of economic specialization,
_ Context, Association, and Provenience
23
pri31057_ch01.pp1-33.qxd
10/15/09
11:18 AM
Page 24
Figure 1.26 An early Neolithic site in northern Germany showing some of the activities and objects associated with the occupation. Clockwise from the top: bow hunting, wild elk and boar, pottery, a fishing spear, an antler axe, domestic sheep and cattle, domestic wheat, wild berries, tree felling, a dugout canoe and paddles, a ceramic lamp, flint blades, a groundstone axe, flint tools, and a fireplace.
wherein certain materials were produced by skilled craftspeople, whereas other items were made in individual households. Regional settlement patterns that are recorded in archaeological surveys can provide a variety of information on the prehistoric use of the landscape. Often, several kinds of sites are found in an area (Figure 1.27). Residential settlements of various sizes and durations are typical targets for investigation. Such sites can vary from camps to villages, towns, or prehistoric cities (Figure 1.28). There are many other kinds of sites. Extraction sites are used for more specific, nonresidential purposes to obtain raw materials or resources, such as quarries for stone or copper and places where animals were killed and butchered. Distinct burial areas, outside settlements, are another kind of site. Cemeteries of inhumation graves, cremation urns, or individual burial mounds and tombs are some other types of sites. Ritual or ceremonial areas may be isolated localities on the prehistoric landscape—Stonehenge, for example.
Interpretation of Archaeological Information
The most difficult thing to predict is not the future, but the past. —Russian proverb
_ _
24
Archaeological information that is recovered from the ground and described and analyzed by specialists does not directly say very much about the past. The analyses may tell us what the items were, what they were made of, how they were used, and how old they are. But the questions that archaeologists seek to answer about the past concern larger concepts: the way of life of prehistoric peoples, the way human societies coped with their physical and social environments, and the way our predecessors viewed their world. Both the questions we ask and the ideas we use to find the answers are at the heart of interpretation in archaeology. The science and the creativity of archaeology lie in bridging that gap between the information we recover and the questions we seek to answer.
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 25
Ancient urban center
City, town, large and small villages
Figure 1.27 A schematic representation of the changes in social systems from small bands to urban societies in prehistory. Note also that settlement patterns change over time and that bigger and more diverse communities emerge.
Towns, large villages, and small villages
Large and small villages
Small villages
Camps
The Science of Archaeology The questions that archaeologists seek to answer are highly varied. Some questions are very specific: How was a flint scraper used? Others are very general: Why did humans domesticate plants and animals? Many questions arise from simple curiosity or common sense—the basic when, where, and what kinds of questions. Other questions—usually the how and the why kind—are much more difficult. Some questions and some ideas come from our knowledge of living peoples. Ethnographies—anthropological descriptions of living or historically known groups of people in different parts of the world—are a prime source of information about human behavior and its variation. Archaeologists often look for comparisons in ethnographies to explain things. For example, knowing that the Northwest Coast Indians in North America lived in longhouses in the early twentieth century helps us make sense of buried structures built centuries earlier in that region. But ethnographic information has limitations, since past human behavior is much more diverse than what has been recorded in the anthropological literature about living peoples. As an extreme example of this situation, there are no ethnographies of Neanderthals. Explaining human behavior in the past and its changes through time is a major goal of archaeology. Big questions about the past often come from our ideas about how things worked. Archaeological theories are bodies of ideas about human behavior in the past. There are also many theories about that past, but many of these tend to be related. There are, in fact, groups of theories that define schools of thought or perspectives in archaeology. The history and theoretical underpinnings of archaeology are beyond the scope of this book; several places to look for more information can be found in the Suggested Readings section at the end of this chapter. We can, however, look at some of the basic aspects of past societies and behavior that theories try to explain.
Archaeology is the search for fact. Not truth. If it’s truth you’re interested in, Doctor Tyree’s philosophy class is right down the hall. So forget any ideas you’ve got about lost cities, exotic travel, and digging up the world. We do not follow maps to buried treasure and “X” never, ever, marks the spot. Seventy percent of all archaeology is done in the library. Research. Reading. —Indiana Jones (1989). Indiana Jones and the Last Crusade. Screenplay by Jeff Boam, story by George Lucas and Menno Meyjes.
ethnography The study of human cultures through firsthand observation.
Context, Association, and Provenience
25
_
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 26
1:40,000,000
1:4,000,000 Maas R. Settlement cells
R.
R. ine Rh
Rhine
Central European Neolithic
Danu b
eR
.
Findspots R
hin
e
R.
1:400 Postholes
House clusters
N
e
M
ch
a rzb
R.
1:400,000
Rubbish pits Paths
1:40,000
Old house sites
Silos
Occupied houses 1:4000
Figure 1.28 A series of views of the distribution and pattern of settlement in the Early Neolithic of central Europe. These maps and plans show the nature of the Linearbandkeramik settlements approximately 7000 years ago.
_ _
26
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 27
Our theories and ideas are basically attempts to explain what took place in the past. The hard part of archaeology is connecting the facts (data) and the ideas (theories) to better understand what happened. The process of asking and trying to answer questions is essentially the process of learning. What makes archaeology science is rigorously testing or evaluating the answers to be confident that they are not wrong. Aspects of Society and Behavior The kinds of questions that archaeologists ask about past societies in general terms involve concepts such as technology, economy, organization, and ideology. Technology is the set of tools, techniques, and knowledge that allows people to convert natural resources into tools, food, clothing, shelter, and other products and equipment they need or want. Technology is the means by which people interact directly with their natural environment. It is also the aspect of past culture that is most easily observed in archaeology. The fragments of the tools that people used in the past, made of durable materials such as stone, ceramic, and metal, are the most common archaeological remains (Figure 1.29). Changes in technology over time provide clear indicators of the development of our nonbiological means of adaptation. Economy is a broad topic that involves how people obtain foods, materials, and goods to sustain their lives. One major aspect of prehistoric economies is subsistence—the activities and materials that people use to feed themselves. Archaeologists use the term subsistence pattern to describe the plants and animals that prehistoric people ate, the activities required to obtain those foods, and the procurement and preparation techniques and implements used to turn those plants and animals into food. The term hunting and gathering describes one general pattern in which wild animals are hunted and wild plants are collected or gathered for subsistence. Agriculture is a subsistence pattern that involves the herding of domesticated animals and the cultivation of domesticated plants. Exchange is an important aspect of economy (Figure 1.30). When artifacts such as stone axes, obsidian knives, metal spearpoints, or certain kinds of food are passed from person to person or group to group, archaeologists talk about “exchange.” One way to study such interaction within and between societies is to look at the distribution of items of exchange. Economic anthropologists distinguish three kinds of exchange: reciprocity, redistribution, and trade. Reciprocal exchange sometimes takes the form of gift-giving, whereby objects of relatively equal value are given to build alliances. Redistribution involves the movement of goods to a central place from which they are portioned out to members of a society. Such a system of redistribution may be used to support an army, or priests, or the pyramid builders of ancient Egypt. Large-scale economic transactions known as trade often involve some sort of market economy and perhaps a monetary standard. Trade takes place in our own economic system today: Objects are imported and exported for the purpose of making a profit. This level of exchange usually involves a highly complex society with professional artisans, regular supplies of raw material, extensive transportation systems, protection of markets and traders against thieves, and enough customers to make the business worthwhile. Archaeologists often examine exchange and interaction through the study of “exotic materials.” The presence of objects and materials that are not available or locally produced in the study area provides immediate evidence of connections and interaction with others. Of greatest use in such investigations are artifacts or materials that come from a single location. Organization refers to the roles and relationships in society and concerns relations between women and men and among different segments of society, such as families, age groups, labor units, or ethnic groups. Organization structures various aspects of society, such as social interaction, economic activity, and political relationships.
Figure 1.29 blade.
An obsidian core and
technology The combination of
knowledge and manufacturing techniques that enables people to convert raw materials into finished products. economy The management and organization of the affairs of a group, community, or establishment to ensure their survival and productivity. organization The arrangements between individuals and groups in human society that structure relationships and activities.
Context, Association, and Provenience
27
_
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 28
Emissary trade
Market Port of trade
Food
Beads Food
bo
un
Reciprocal exchange
Beads
Food Beads
dar
Reciprocal exchange
Direct access
y Stone for tools Down-the-line trade
Figure 1.30 The exchange of goods can take many paths in human societies. This diagram shows some of the methods for exchange in prehistoric economies. Such patterns reflect both economy and organization.
egalitarian A term that refers to societies lacking clearly defined status differences between individuals, except for those due to sex, age, or skill. rank A relationship of inequality between members of society in which status is determined by kinship relations of birth order and lineage.
_ _
class A relationship of inequality between members of society in which status is determined by membership in a level or class.
28
Down-the-line trade
Kinship and marriage systems, lineage, rank, and class are important aspects of social organization and a means of structuring social relationships. Kinship defines the relationship between individual members in society on the basis of their family relationships. Grandmother, brother, uncle, and cousin are terms that relate us to other people through kinship. Marriage systems tie unrelated individuals together through sanctioned kinship; rules for these relationships are carefully defined in society. Lineages provide a means for calculating one’s relationships through lines of ancestry. Such genealogies are a way to extend relationships and determine membership in a group. Members of the same lineage often work as a corporate group. Rank and class distinguish individuals and groups of people within society. Many societies of hunter-gatherers are described as egalitarian, with essentially equal relations between all members of the group. Many agricultural societies are larger and exhibit distinctive groups within the society that are defined by inherited status differences. Higher status (resulting from prestige, wealth, and/or power) characterizes elite and privileged groups in a society. Rank and class are means of defining such status groups. Rank refers to inherited positions in societies in which everyone is ranked by status relative to all other people. The firstborn of the highest-ranked group is the highest position in such a society. In ranked societies, each individual has a unique place in the order of relationships. Class societies are structured by distinctions between groups, or classes, of people that define levels, or strata, in society. Class is also usually inherited but defines large groups of individuals and may determine one’s job, location of residence, marriage opportunities, and financial status. India under the caste system was an extreme example of a society structured by class. The economic activities of prehistoric peoples were organized in various ways. A fundamental mechanism for the organization of tasks is the division of labor. Separate groups or segments of society undertake different activities as part of the economic process. A basic example is seen in many groups of huntergatherers in which the division of labor is by sex; males are primarily hunters, and females are primarily gatherers. Both groups contribute foodstuffs to the subsistence economy of the group. Agricultural societies also see economic orga-
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 29
nization along gender lines, but the household becomes an important component of production for food and other necessary materials. Production becomes more specialized over time: Entire communities may be involved in the production of specific items, or specialist groups of producers—craft specialists—such as potters, metalsmiths, and beadmakers, may emerge. Production units can assume more formal structures, such as guilds or unions, in larger, more complex societies. In a general sense, political organization is a reflection of the increasing complexity that is witnessed in human society over time. As societies became larger, organizational changes resulted in closer integration and more linear decision making. One of the most significant changes in organization was the shift from egalitarian to hierarchical structures that often followed the origins of agriculture. Hierarchical organizations have one or more levels of control above the majority of the people in the society (Figure 1.31). These higher levels are seen in elite classes or ranks that control much of the wealth, power, and decision making in society. One way to imagine such a hierarchy is to recall the nature of military organization and the chain of command from generals to lieutenants to privates. Government also operates in a hierarchical manner, from local representatives to municipal government, state government, and federal government. The sphere of control and decision making varies with the level in the hierarchy. The municipal government repairs local roads; the federal government builds an interstate highway system. There are several ways to describe or characterize such hierarchies in human society (Figure 1.32). The concepts of bands, tribes, chiefdoms, and states are often used to distinguish different kinds of political organizations. Bands and
Figure 1.31 Pyramid of the social layers in a medieval European population. Note also the symbols that mark positions in society—for example, the soldiers’ helmets, the monks’ haircuts, and the king’s crown.
craft specialists (or craft specialization) Individuals involved in part-
or full-time activities devoted to the production of a specific class of goods, often highly valued. hierarchical A term referring to soci-
eties that have a graded order of inequality in ranks, statuses, or decision makers.
_ Context, Association, and Provenience
29
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 30
Hunter-gatherer bands
Figure 1.32 Types of societies and the appearance of various institutions. This chart documents the emergence of many of the major institutions in human society on the path from small bands to more complex civilizations.
Hierarchical states TYPE OF SOCIETY
Taxation Military draft Bureaucracy Codified law Kingship Stratification Full-time craft specialization Separate marriage networks for elites and commoners
Some institutions, in approximate order of appearance
Hereditary leadership Central accumulation Ranked descent groups Regularly scheduled calendric rituals Pan-tribal associations Unranked descent groups Reciprocal or barter economy Unscheduled ad hoc ritual Ephemeral, informal leadership Egalitarian status Local group autonomy
ideology A conceptual framework by which people structure their ideas about the order of the universe, their place in that universe, and their relationships among themselves and with objects and other forms of life around them. cosmology The worldview of a group
_ _
or society, encompassing their understanding of the universe, their origins and existence, and nature.
30
tribes describe relatively small societies of hunter-gatherers or farmers in which relationships are generally egalitarian and decision making is consensual. Power and property are distributed among all the members of the population. Status is earned through achievements and is ephemeral, held only by the individual who gained it. Chiefdoms and states are larger, often territorial, societies in which relationships are unequal and organization is hierarchical. Status is hereditary and assigned or ascribed by birth order or class affiliation. One of the more apparent (though not always present) trends in the organization of human society is an increase in complexity over time. Complexity refers to more different units in society and more integration between those units. More units are a result of social, economic, and ideological specialization. Such differentiation is reflected in the distinctions between villages, towns, and cities that began to appear with chiefdoms and states. More integration is a result of hierarchical organization and the emergence of ranked or stratified groups within society whereby power and decision making are in the hands of a few. Some of the changes in social, economic, and political organization from bands to states are summarized in Figure 1.32. Ideology refers to the means by which people structure their ideas about the universe, their own place in that universe, and their relationships with one another and with things and other beings around them. Ideologies are shared but also vary within and among groups of people. Ideology is the way that people view and understand their world. That view often affects many things they do. Ideology is reflected in the clothes we wear, the food we eat, and the places in which we live. Ideology encompasses the norms, values, and beliefs held by a society. Ideology is reflected in cosmology—explanations of the origins of the universe, of life, and of society. Roman cosmology invoked the twins Romulus and Remus, mythical beings raised by a she-wolf, as the founders of Rome.
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 31
Symbols and styles often are expressions of ideology and identity. School mascots, corporate logos, religious icons, national flags, and certain faces incorporate and display a wide range of concepts and ideas. Think of the importance attributed by thousands or millions of people today to the leprechaun of the University of Notre Dame, the circled star of a Mercedes, a statue of Buddha, the American flag, or a picture of Fidel Castro. Symbols, styles, and ideologies often are expressed in art. Among huntergatherers, decoration often appears to have been individualistic, marking ownership or life events. Distinctive distribution of styles can be observed that probably mark the range of particular groups. For example, there were different zones of spearpoint styles in the late Paleolithic in northern Europe. From the beginning, there were distinctive styles of art. The first art appeared toward the end of the Paleolithic, about 30,000 years ago, in Europe, Africa, and Australia (Chapter 4, p. 105) and reflected the worldview of the people who produced it. Most of the cave paintings from the European Paleolithic are of large game animals, the major prey of these human groups—reindeer, bison, wild cows, and many other species. There are very few humans, plants, or scenes depicted. The cave and rock art of prehistory are of two major types. Pictographs were made by the application of pigment to rock surfaces (such as cave paintings); petroglyphs were made by removing the outer surface of a rock by carving or hammering (Figure 1.33). Ideology is particularly clear in the art of larger civilizations, often in the form of propaganda. Distinctive motifs, like modern corporate icons, can be found everywhere the power of a particular political entity extends. This pattern can be seen in the double crown of the Egyptian pharaoh, the were-jaguar motif of the Olmec of Mexico (pp. 335–336), and many other contexts. Much of the art and decoration we know from past civilizations served both aesthetic and political purposes. Ideology is frequently expressed in ceremony and pageant surrounding important rites of passage through life: birth, adulthood, marriage, and death. Ideological norms and rules are usually stored and maintained in the older generation. Ideology is often embodied in specialists who maintain ritual knowledge and direct the ceremonies and activities that keep such ideology active and pertinent. In egalitarian societies, such individuals are known as witches and shamans—specialists in ritual and healing, seers of the future. In hierarchical societies, such specialists are found in powerful groups such as priesthoods, political organizations, and other sodalities. These components of human society—technology, economy, organization, and ideology—are closely interrelated in prehistoric materials. A single artifact or object may contain aspects of each. A type of knife found exclusively in women’s graves may hold information on the manufacture of tools, on the nature of women’s work, on the distinction between sexes in the society, and on ideas about death. Technology, economy, organization, and ideology thus are different but related dimensions of past cultures and of human life and an important focus of archaeological investigations.
Figure 1.33 A rock carving of a Viking ship from Norway. Red has been used to highlight the carving.
How does art function in the world today?
y
What kinds of sodalities are found in modern society?
pictograph A written or painted symbol that more or less portrays the represented object. petroglyph A drawing that has been carved into rock. shaman An anthropological term for a spiritualist, curer, or seer. sodality An alliance or association among some members of a society, often based on age and sex, with a specific function.
Context, Association, and Provenience
_
31
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 32
Images and Ideas The Basics of Archaeology
T
his chapter has provided an introduction to the basic themes of archaeology— the study of past human behavior. Important concepts for thinking about the past are time and change. Time is so vast that it is usually subdivided to make it more comprehensible, just as hours and minutes divide the day. Archaeologists utilize two systems for time, one geological and one archaeological. Geologists study the entire history of our planet: 4.6 billion years. They divide that time into major eras based primarily on changes in life-forms: Precambrian, Paleozoic, Mesozoic, and Cenozoic. The Cenozoic, the important one for humans, is divided into a series of epochs. The last four are the Miocene (when the first human ancestors separated from the apes), the Pliocene (when varieties of human fossil species emerge), the Pleistocene (the Ice Age, when humans leave Africa), and the Holocene (the past 11,000 years). Archaeological time essentially covers the period of humans on earth, from the end of the Miocene. In fact, archaeological time really begins with the first artifacts, about 2.6 million years ago. The divisions in archaeological time are based on changes in the artifacts that people made in the past. In general, the important artifacts for chronology are made of stone, ceramic, or metal, hard materials that survive in the ground. Major divisions of archaeological time are based on the raw materials used for tools: stone, bronze, and iron. The Stone Age is divided into the Paleolithic, the Mesolithic, and the Neolithic, and these are further subdivided in many places. In the context of time, archaeologists deal with change in human behavior. Such changes can be related to biology and/or culture. Biological changes are described as evolution. The earliest recognizable humans evolved from an apelike ancestor and became more modern over time (Figure 1.34). The primary mechanism driving the biological changes in humans during the Miocene and the Pliocene was natural selection, the engine of evolution. The first artifacts— very simple, sharpened edges of stone—date to the end of the Pliocene. This is the initial evidence of culture. Culture is that uniquely human mode of adapting to the world around us. Culture is a buffer between us and the harshness of nature. Culture is learned behavior; culture is tools, information, organization, and action. Culture is stone-cutting implements; culture is building fires to stay
< ( : < V : y < p : W
Figure 1.34 A chimpanzee using a simple tool for immediate reward. An adult chimp uses a twig to extract insects from a hole in a tree. Humans make tools in order to make other tools for later use, a unique distinction.
_ _
32
Chapter 1 Principles of Archaeology
pri31057_ch01.pp1-33.qxd
10/1/09
7:46 AM
Page 33
warm; culture is an incest taboo. Cultural development became the main means of human adaptation during the Pleistocene period, replacing biology as our primary mechanism for coping with stress and change. Archaeologists use theory and method to obtain information about human culture and behavior in the past and try to answer basic questions such as what, when, where, who, how, and why. Archaeological investigations generally follow a three-stage process involving discovery, analysis, and interpretation. Survey and excavation are the discovery components of fieldwork that are intended to find and recover artifacts, sites, and information. Analysis is the long and involved study of archaeological materials, describing, identifying, quantifying, compiling large quantities of data, and extracting answers to some of the easier of those basic questions. Interpretation is the difficult process of trying to make sense of the data and to understand what happened in the past. Interpretation usually involves the harder questions of how and why things happened and changed in the past. These big questions include the following: Why did we start to walk on two legs and become human more than 6 m.y.a.? What was responsible for the creative explosion in human behavior that was witnessed late in the Pleistocene? Why did we begin to domesticate plants and animals of the earth at the end of the Pleistocene? Why did inequality replace principles of egalitarianism? How did the first cities arise and operate? Why did empires die? The investigation of those questions and others is the subject of the remaining chapters in Images of the Past.
p
It’s not what you find, it’s what you find out. —David Hurst-Thomas (1989)
DISCUSSION QUESTIONS 1. What are the two important chronological systems used in archaeology? Can you think of other methods for calculating time? 2. How does biological evolution work? How does one species evolve into another? 3. What are some of the main methods of archaeological fieldwork? 4. What are some of the kinds of sites that may be found during archaeological fieldwork? 5. What are some of the directions in which the analysis of archaeological materials can go?
z
w w w. m h h e . c o m / p r i c e i p 6 e
For more review material and study questions, see the self-quizzes on your online learning center.
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e.
Bahn, P., and C. Renfrew. 1996. The Cambridge illustrated history of archaeology. Cambridge: Cambridge University Press. An exceptionally well-documented depiction of the development of archaeology. Brothwell, D., and A. M. Pollard, eds. 2001. Handbook of archaeological sciences. London: Wiley Europe. A comprehensive survey of the various aspects of archaeometry and archaeological science. Collis, J. 2001. Digging up the past: An introduction to archaeological excavation. Stroud, UK: Sutton. Drewett, P. 1999. Field archaeology. An introduction. London: Routledge. An up-to-date manual on field archaeology and its primary methods. Greene, K. 2002. Archaeology: An introduction, 4th ed. London: Routledge. An excellent how-to introduction to archaeology. Hodder, I., ed. 2001. Archaeological theory today. Cambridge: Polity Press. A series of papers on the present state of archaeological thinking assembled by a leading practitioner.
Johnson, M. 1999. Archaeological theory. Oxford: Blackwell. A discussion of current thinking in archaeology with comment and critique. Marcus, J., 2008. The archaeology of social evolution. Annual Review of Anthropology 37:251–266. An up-todate discussion of the evidence for social evolution in the past. Price, T. D. 2007. Principles of archaeology. New York: McGraw-Hill. A new consideration of the major methods and theories of archaeology. Renfrew, C., and P. Bahn. 1998. Archaeology: Theories, methods, and practice. London: Thames & Hudson. A popular textbook emphasizing how archaeologists learn what they know about the past. Sutton, M. Q., and B. S. Arkush. 1998. Archaeological laboratory methods. Dubuque, IA: Kendall/Hunt. A guide to laboratory work in archaeology involving identification, conservation, and measurement.
Suggested Readings
33
_
pri31057_ch02.pp34-71.qxd
10/1/09
C
7:48 AM
H
Page 34
A
P
2 T
E
R
T
W
O
_ _ Figure 2.1 The 3.75-million-year-old footprints discovered in East Africa at the site of Laetoli, Tanzania.
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 35
< The First Humans ( Introduction : T < V : y < p : W The Dawn of Humanity
he 3.75-million-year-old footprints found at the site of Laetoli in East Africa (Figure 2.1) are the most dramatic and important evidence we have of early humans as two-footed animals. Becoming bipedal appears to have been the first step we took in separating from our apelike ancestors. Our evolutionary path to being human is the subject of this chapter. To comprehend the significance of this transformation, it is essential that we understand our place among the other animals. Zoologists classify the members of the animal kingdom according to their similarities and differences. We are animals because we move and eat with a mouth, we are vertebrates because of our backbone, and we are mammals because we have warm blood and breastfeed our offspring. We are primates because we have grasping hands, flexible limbs, and a highly developed sense of vision, which we share with the other members of the primate order: lemurs, tarsiers, monkeys, and apes. We are members of the family Hominoidea, the taxonomic group that includes apes and humans, because of the shape of our teeth, the absence of a tail, and our swinging arms. As humans, we share with chimpanzees and gorillas a common ancestor that lived sometime during the past 5–10 million years. Present evidence suggests the following scenario for primate evolution during the Cenozoic. The first primates on Earth existed about 65 m.y.a. at the beginning of the Cenozoic, when the air temperature was warm and extensive tropical forests covered much of the land surface. These early primates began as treedwelling insect-eaters. They had adaptive characteristics such as stereoscopic color vision, which provided depth perception and enhanced their ability to move from branch to branch and to spot insects, and a grasping ability, so they could hold onto branches and grab bugs. This heritage has provided us with an extraordinary visual ability and large centers in the brain to process the enormous volume of information absorbed by the eyes. Along with the ability to hold and manipulate objects with dexterity came other changes. Arms and shoulders became more flexible for swinging in the trees, and internal organs and bones evolved toward a more vertical arrangement. This is evolution in action. The earliest Hominoidea appeared approximately 25 m.y.a. Apes are generally distinguished from monkeys and other primates by larger size, distinctive teeth, greater sociability, the absence of a tail, and a reduced sense of smell. From one of these early apes, a new group of animals, known as the dryopithecines, emerged during the Miocene epoch some 17–12 m.y.a. These creatures had several features, known primarily from the fossil teeth that have survived, suggesting that they were the probable ancestors of both living apes and humans. Dryopithecines were apparently very successful in their arboreal adaptation, ranging over much of Africa, Asia, and Europe. During this time, the earth changed dramatically. Increased geological activity in the earth’s crust created new mountain ranges.
35
CHAPTER OUTLINE
Introduction The Dawn
of Humanity 35
CONCEPT The Family Tree SITE Hadar
39
44
CONCEPT Dating
Methods 48
SITE Laetoli
50
SITE Swartkrans
52
CONCEPT Hunters or
Scavengers? 55
SITE Olduvai
56
CONCEPT The Leakey
Family 61
CONCEPT The First Tools
63
Images and Ideas
Bones, Stones, and Human Behavior 66
primate The order of animals that includes lemurs, tarsiers, monkeys, apes, and humans.
Hominoidea The taxonomic group
(family) that includes the human and ape members of the primates, both fossil and modern forms.
dryopithecine The generic term for
the Miocene fossil ancestor of both the living apes and modern humans, found in Africa, Asia, and Europe.
_
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 36
Figure 2.2 The skulls of Australopithecus afarensis (left), Homo erectus (center), and Homo sapiens sapiens (right).
Man is only one of the earth’s “manifold creatures” and he cannot understand his own nature or seek wisely to guide his destiny without taking account of the whole pattern of life. —George Gaylord Simpson (1967)
w w w. m h h e . c o m / p r i c e i p 6 e
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
y
Are humans really so different from other animals?
Figure 2.3 An artist’s reconstruction of Kenyapithecus.
_ _
36
Chapter 2 The First Humans
Between 9 and 4 m.y.a., the convergence of the Indian and Eurasian continental plates gave rise to the Himalayan Mountains and Tibetan Plateau. These massive new elevations rerouted the weather, and the climates from East Africa to East Asia became drier and more seasonal. Widespread tropical forests began to shrink, taken over by expanding grasslands and savannas. It is very possible that these events pushed some forest-dwelling apes in Africa toward the open savanna and along the human line (Figure 2.2). At some point during the later Miocene epoch, after 10 m.y.a., one of these African primate species took the path toward humanness, as seen in the evidence for more upright posture and smaller canine teeth. The fossil record from this time period is very scanty. A fossil discovery in the Sumburu Hills of Kenya may be this creature (Figure 2.3). A fragment of an upper jaw, dating to 9–8 m.y.a., has characteristics of a generalized chimpanzee-gorilla-human ancestor. Genetic and molecular evidence indicates that early humans were most closely related to chimpanzees and that we began to diverge from the chimpanzee lineage between 6 and 5 m.y.a. Recent fossil evidence from Central and East Africa suggests that this date may be even earlier, before 6 m.y.a. Between 6 and 2 m.y.a., a variety of our early relatives made their appearance in Africa. But what does it mean to be human? What makes us distinct from other species of animals? We are human because we have a skeleton designed for upright walking. We are human because we have grasping hands with opposable thumbs, capable of both strength and precision movement; we are also human because we have lost the grasping, opposable toes of the other apes. We are human because we have small, flat teeth and lack the large, slashing canines of other primates. We have a pronounced nose compared with that of apes and a face that sits beneath our brain case rather than in front of it. We are human because we lack fur and have more sweat glands than hair follicles. We have a conspicuous penis and breasts, and we have sex face-to-face and almost constantly compared with many other animals. We care for our young over a lengthy period of infancy and childhood. We are human because we make and use tools to alter our environment and make our lives more secure and comfortable. We are human because we
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 37
Cranium Mandible
Lumbar curve Ribs Large lumbar vertebrae
Short lumbar region of spine
Short, broad ilium
Powerful arms
Ilium Pubis Ischium
Pelvic bone
Long, strong thumb
Long, narrow ilium
Broad sacrum
Long, powerful fingers, weak thumb
Long, mobile toes Mobile ankle Mobile big toe
Long, slender femur Knee “locks” in full extension
Ankle stable, little rotation Short toes
Double-arched foot Strong heel
Figure 2.4 Comparative skeletal anatomy of the chimpanzee and the modern human.
often act according to reason rather than on instinct. We have a large brain relative to our body size, and we have enhanced intelligence, as well as a complex repertoire of behaviors known as culture. We are human because we speak a language full of meaning and metaphor. Many features, then, define us as human (Figure 2.4). But what characteristics can be identified in the fragmentary fossil bones and stone tools of our earliest ancestors? The markers of early humanness must be present in materials that can survive thousands and millions of years of exposure to the elements. Of all the characteristics of being human, the ones that can be found most readily in the fossil record are upright posture, larger brains, and tools. The major questions in human evolution, then, concern when, where, and why these distinctive characteristics appeared. What of us is preserved in the layers of geological time? Paleoanthropology, the study of early human evolution, attempts to answer those and other questions, using evidence from fossils and artifacts. For this information, we must turn to Africa; our oldest ancestors are known only from that continent. Some of the best evidence comes from sites such as Hadar, Laetoli, Swartkrans, and Olduvai (Figure 2.5). We will visit these places in the following pages.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on primate evolution, see the Internet exercises on your online learning center.
paleoanthropology The branch of anthropology that combines archaeology and physical anthropology to study the biological and behavioral remains of the early hominins. fossil The mineralized bone of an
extinct animal.
_ Introduction The Dawn of Humanity
37
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 38
Figure 2.5 Location of and timeline for early hominins in Africa.
Lake Chad Basin
In each great region of the world the living mammals are closely related to the extinct species of the same region. It is, therefore, probable that Africa was formerly inhabited by extinct apes closely allied to the gorilla and chimpanzee; and as these two species are now man’s nearest allies, it is somewhat more probable that our early progenitors lived on the African continent than elsewhere. —Charles Darwin (1871)
Hadar Middle Awash Omo Lake Turkana Tugen Hills Lake Victoria Olduvai Laetoli
ATLANTIC OCEAN
Equator
INDIAN OCEAN
Location of sites mentioned in this chapter Approximate distribution of hominins in the Pleiocine Makapansgat Taung Lake Chad Basin
Swartkrans Sterkfontein Kromdraai
Sahelanthropus tchadensis
Tugen Hills
Orrorin tugenensis Middle Awash
Ardipithecus ramidus
Lake Turkana
Australopithecus anamensis
Hadar Laetoli
Australopithecus afarensis Australopithecus afarensis Homo erectus South Africa
Paranthropus robustus Australopithecus africanus Homo habilis Olduvai
Paranthropus boisei Homo habilis Paranthropus boisei
Koobi Fora
Australopithecus africanus
_ _
7
38
6
Chapter 2 The First Humans
5
4 Millions of years ago
3
2
1
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 39
Concept
<
The Family Tree The evidence and interpretation of our earliest biology
Human evolution—the changes in the skeleton and biology of our species over the past 7 million years—is a fascinating subject, but one for which there is only sparse evidence. In almost every instance, the fossil remains of our earliest hominin ancestors are very fragmentary, poorly preserved, and disturbed by natural forces—time and nature have taken their toll. Since the initial work of the Swedish botanist Carolus Linnaeus during the mid-1700s, scientists have classified newly discovered members of the plant and animal kingdoms according to a system that organizes them into species, genus, family, order, class, phylum, and kingdom, from most specific to most general—a family tree of life. Modern humans are members of the family Homininae of the genus Homo and the species sapiens. Determining the genus and species of the fossil bones of early humans is very difficult. All the fragmentary early fossil finds represent only a few parts of a few hundred individuals. Determining the age of fossils is difficult, and questions remain about whether the species existed contemporaneously or sequentially. Not surprisingly, there are controversies over what to call these first hominin forms and how to identify them. As Richard Klein, of Stanford University, has noted, paleoanthropology is more like a court of law than a physics laboratory. It sometimes seems that whenever a new fragment is discovered, we have to reassess and even redraw our entire family tree. New fossils that modify current ideas are found almost every year. Disputes rage over the designation of species, the age of fossils, and the line of human ancestry. Even the words to use to describe the category of early human ancestors are controversial. New fossil finds have
Central chimpanzees
Eastern chimpanzees
Modern humans
Nigerian chimpanzees Western chimpanzees
Eastern gorillas
6 m.y.a.
Bornean orangutans
Mountain gorillas
Bonobos Sumatran orangutans 12 m.y.a.
Western gorillas
25 m.y.a. Common ancestor of great apes
Figure 2.6 An evolutionary tree for the great apes based on modern genetic studies. The length of lines shows genetic distance from one group to another. The diagram shows that our closest relatives are the chimpanzees and the bonobos and that we are not very far apart compared with some of the other species. Three estimates for the time elapsed since separation are shown on the diagram. Fossil evidence suggests that the separation of humans and an ape relative may be somewhat earlier than 6 m.y.a. bonobo A small species of chimpanzee, closely related to humans.
forced a reconsideration of the terms because it becomes harder to distinguish the first humans from their closest relatives among the apes. The term hominoid refers to all present and past apes and humans. The word hominid has been used for many years as the
hominoid A descriptive term for any
human or ape, past or present, characterized by teeth shape, the absence of a tail, and swinging arms. hominid An obsolete term that refers to the human members of the primates, both fossil and modern forms.
Concept The Family Tree
39
_
pri31057_ch02.pp34-71.qxd
Gibbon
10/1/09
7:48 AM
Orangutan
Page 40
Gorilla
Chimpanzee
Human
2.0 4.4 5.7 ± 1.5 6.2 ± 1.8
Millions of years ago 11.6 ± 1.5
Sexual dimorphism Thick enamel
14.4 ± 2.0
Thin enamel
Bipedalism
Figure 2.7 A diagram showing the speciation of the great apes with chronology and some distinguishing characteristics. Humans, for example, exhibit bipedalism, thick enamel, and little sexual dimorphism. The red X indicates reduced sexual dimorphism.
An ape-brained and small-canined creature, with dental enamel of unknown thickness. Large if male but smaller if female. May be spotted climbing adeptly in trees or walking bipedally on the ground. Last seen in Africa between 5 and 7 million years ago. —Pat Shipman (2002), describing the earliest humans
hominin A current term that refers to the human, chimp, and gorilla members of the primates, both fossil and modern forms.
_ _
sexual dimorphism A difference in size between the male and female members of a species.
40
Chapter 2 The First Humans
generic term for present and past humans. However, a new term, hominin, is now being used in place of hominid. Genetic studies have shown that not all apes descended from a common ancestor (Figure 2.6)—that chimps and gorillas share a more recent ancestor with humans than they do with the orangutan, for example. That means that, on the strict taxonomic level, chimps and gorillas are hominins. The term hominins, then, is used to describe those species clearly in the line of human evolution, not other apes. All hominins are hominids, but not all hominids are hominins. Confusing, isn’t it? Recent years have seen major changes in the field of paleoanthropology—new discoveries, new dates, new species, and new disagreements. A flurry of fossil finds has spurred these changes and once again rewritten our understanding of the evolution of our earliest ancestors. These discoveries have come from an important new locale in Chad, in central Africa, and from the usual places in East and South Africa. Features of the fossil material
such as thickness of tooth enamel, evidence for bipedalism and tree climbing, and size differences between the sexes— sexual dimorphism—play an important role in the discussion (Figure 2.7). In Chad, a new, very old, and controversial fossil species has been found in the blowing sands of the Late Miocene deposits in the Lake Chad basin. This new form is designated as Sahelanthropus tchadensis. Although it combines ape and human characteristics, a flat face and “habitual bipedalism” distinguish the Chad specimen as a human ancestor. Habitual bipedalism means the species normally moves on two feet (e.g., humans). Facultative bipedalism means that the species is able to move on two feet (e.g., chimpanzees and gorillas). Dates on this bipedal individual lie between 7 and 6 m.y.a., making it our oldest known ancestor. In East Africa, a spate of finds documents the diversity of our early ancestors. In the Middle Awash area of Ethiopia, seventeen fossils, including teeth, skull, and arm bones, were found by a research team directed by Tim White of the University of California, Berkeley. These fossils lie directly under a volcanic deposit dating to 4.4 m.y.a. Designated Ardipithecus ramidus, the species exhibits a combination of human and chimpanzeelike features. Since the initial discovery, leg bones have been found that suggest that this earliest ancestor likely walked on two legs. Meave Leakey and her colleagues have named a new genus and species, Kenyapithecus platyops, which has a flat, humanlike face but an ape-size brain. This fossil probably dates to 3.5 m.y.a. An even earlier and more important find, from the Tugen Hills of Kenya, are fossils—named Orrorin tugenensis—that date to 6 m.y.a. A CAT scan of the fossil femur (thighbone) indicated habitual bipedalism. Like those of Ardipithecus, the arm bones suggest tree-climbing adaptations, but the two species differ in enamel thickness. Orrorin has thick enamel like a human’s, whereas Ardipithecus has thin enamel more closely resembling that of other chimps and gorillas.
10/1/09
7:48 AM
Page 41
In South Africa, the big news is dating. New techniques have redated the deposits at several sites where varieties of australopithecines have been found and have pushed back the dates to almost 4 million years, twice as long ago as previously believed. These fossils are now easily as old as their counterparts in East Africa. These recent fossil finds have pushed the antiquity of humans and their ancestors back to the end of the Pliocene into the Late Miocene, the oldest more than 7 m.y.a., aging our presence on the planet by several million years (Figure 2.8). The new finds have turned the tree of human evolution into more of a bush with a number of branches at the bottom as well. In addition, the finds have provided important new information on habitat, diet, and posture. The older fossil forms appear to have lived in forested environments, in contrast to the more open savanna that is thought to have been the habitat of the australopithecines. It now seems that bipedalism first appeared in the context of the forest rather than the plain. Changes in diet are apparent in the fossil teeth. Relatively small molar size in chimps, Ardipithecus, and Orrorin indicates a diet primarily of fruit and vegetation. Homo erectus and Homo sapiens also have small molars relative to body size. Tooth enamel is thin in chimps, medium in Ardipithecus, and thick in Orrorin, Australopithecus, and Homo. Canine teeth in chimps and Orrorin are large, sharp, and V-shaped in cross section; canines are small, more rounded, and diamond shaped in Ardipithecus and later forms. The combination of thick enamel, large molars, and smaller canines first seen in the australopithecines is thought to show a change in diet from fruits and leaves to more roots, tubers, insects, and other small animals. Microscopic analysis of wear patterns on fossil teeth by Alan Walker, of Pennsylvania State University, indicates that the tooth enamel among early hominins more closely resembles that of herbivores than that of carnivores (Figure 2.9). Examination of the
0
Homo sapiens
Homo neanderthalensis
Chimpanzees
Homo erectus
Paranthropus robustus
1 Homo habilis
2 Millions of years ago
pri31057_ch02.pp34-71.qxd
3 4
Kenyapithecus platyops Australopithecus anamensis
Australopithecus africanus Australopithecus afarensis
5 6 7 8
Paranthropus boisei
Ardipithecus ramidus
Large brain, small teeth, obligate bipedalism Orrorin tugenensis
Small brain, very large teeth, facultative bipedalism Small brain, large teeth, facultative bipedalism Small brain, small teeth, quadrupedal Not enough evidence
anatomy of the wrist, shoulder, pelvis, and thigh of the early australopithecines indicates a pattern of movement, or locomotion, different from that of both the modern apes and humans. Henry McHenry, of the University of California–Davis, noting the curvature visible in hand and foot bones, concludes that these creatures must have spent some time in the trees. Sexual dimorphism was greatly reduced in Homo erectus and may reflect the emergence of monogamous mating systems, in which males and females each have a single mate for long periods of time. Australopithecus anamensis, a transitional form between A. ramidus and A. afarensis, has recently been found at Lake Turkana, Kenya, dating to 4 m.y.a. Sometime around 3.9 m.y.a., A. anamensis evolved into Australopithecus afarensis, well known from Hadar, Laetoli, and elsewhere in East Africa (Table 2.1). A. afarensis exhibits more humanlike teeth and unquestionably walked upright, as seen in the footprints at Laetoli and in the fossil bones themselves. The next series of fossil finds comes from a period generally referred to as the Plio-Pleistocene. This is a combined term for the late Pliocene
Sahelanthropus tchadensis
Figure 2.8 Hominin evolution over the past 7 million years. The diagram shows only some of the major species and their major characteristics: brain size, tooth size, and bipedalism. (Reprinted by permission from Macmillan Publishers Ltd: Nature, “Paleoanthropology: Hominid revelations from Chad” by Bernard Wood, Vol. 148, p. 134. Copyright 2002.)
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the various hominin species, see the Internet exercises on your online learning center.
What advantages could bipedalism have provided for our ape ancestors?
y
australopithecine The generic term for the various species of the genus Australopithecus, including A. ramidus, A. afarensis, and A. africanus. locomotion A method of animal movement, such as bipedalism. Plio-Pleistocene A generic term for
the period of the Pliocene and early Pleistocene, used to describe the age of fossil finds, approximately 3–1 m.y.a.
Concept The Family Tree
41
_
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Figure 2.9 Electron microscope photos of tooth wear on early hominins. The scratched surface from Australopithecus africanus (left) contrasts with the rough and irregular surface of the enamel of Paranthropus (right), thought to have eaten harder gritty foods such as roots and nuts.
Stunning new fossils of hominins that lived three to four million years ago. . . . During this time, we’re dealing with a warmer, wetter Africa that it seems was spawning hominins from the shores of Lake Chad to the caves of Sterkfontein. —Philip Tobias (2002)
Paranthropus Genus of early homi-
_ _
nins, contemporary with Australopithecus, that includes boisei and robustus as species.
42
Chapter 2 The First Humans
Page 42
and early Pleistocene, approximately 3–1 m.y.a. The generally accepted picture is that sometime between 3 and 2.5 m.y.a., A. afarensis split into two separate lineages. One of those lineages continued as the australopithecines (a generic term for various forms of Australopithecus). This line included both gracile and robust forms. The gracile form, with smaller teeth, skull, and body size, known as Australopithecus africanus, appeared shortly after 3 m.y.a. The robust forms, with big teeth and heavy jaws for chewing plant foods, are designated as genus Paranthropus. Several species have been identified, including aethiopithecus, robustus, and boisei. The robust forms have been found in both East and South Africa and eventually became extinct around 1 m.y.a. The other lineage led to Homo habilis, the first members of our own genus. The earliest H. habilis is known from around 2.5 m.y.a. and is recognized by a clear increase in brain size. (The first H. habilis is very close in time to the earliest known stone tools; see “The First Tools,” p. 63.) The 1986 discovery by Johanson and White of over 300 pieces of a skeleton in beds at Olduvai Gorge dating to 1.8 m.y.a. has filled in part of the picture of Homo habilis. For the first time, there were enough fragments of the arms and legs of an H. habilis creature to provide an indication of height and the proportions of the limbs. Surprisingly, this fe-
male H. habilis was less than 1 m (about 3 ft) tall and had very long arms, similar to Lucy and other australopithecines. Such evidence suggests that (1) H. habilis may still have been spending part of its life in the trees, (2) sexual dimorphism was still very pronounced, and (3) major changes in behavior and habitat of the early hominins may have taken place in the period between 2 and 1.5 m.y.a. Thus, it appears that Homo habilis “represents a mosaic of primitive and derived features, indicating an early hominin which walked bipedally . . . but also retained the generalized hominoid capacity to climb trees” (Susman and Stern, 1982, p. 931). Richard Leakey, Alan Walker, and a few others suggest a different scenario. Louis Leakey always argued that the genus Homo had its roots deep in the Pliocene, and he eventually discovered several early Homo specimens at Olduvai and elsewhere. This view is maintained by his son Richard and others who would push the evolutionary split from a common ancestor of the Australopithecus and Homo lines much further back in time, perhaps near the beginning of the Pliocene, around 6 m.y.a. They imagine that two or more different australopithecine groups (P. robustus and A. africanus) and one line of Homo (H. habilis) evolved at this time. The next stage in our family tree is relatively straightforward and uncon-
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 43
TABLE 2.1 Major Characteristics of the Plio/Pleistocene Hominins Australopithecus Australopithecus Australopithecus Australopithecus ramidus anamensis afarensis africanus
Paranthropus robustus
Paranthropus Homo boisei habilis
Dates Sites
4.5–4.3 m.y.a. Middle Awash
4.3–4.0 m.y.a. Lake Turkana
4.2–2.8 m.y.a. Hadar Omo Laetoli
2.2–1.5 m.y.a. Kromdraai Swartkrans
2.2–1 m.y.a. Olduvai Lake Turkana Omo
Cranial Capacity
Unknown (400–450 cc?)
Unknown
Size
100 lb?
5' 110 lb (么) 4'3" 70 lb (乆)
380–500 cc; average = 440 cc 3'6" 50 lb (乆) (么 to 100 lb?)
520 cc (based on one specimen) 5'+ 150 lb
500–530 cc; average = 515 cc 5'+ 150 lb
Skull
Large pointed canines, small molar crowns, thinner enamel; foramen magnum forward
Teeth and jaw hominin, but some similarities to chimpanzee
Heavy jaws, small canines and front teeth, large back teeth, definite crest
Very large jaws, very large back teeth, large crest
Hands and feet more like modern humans’, retention of long arms
—
Postcranial Arm bones Joints on leg Skeleton with character- bones indicate istics intermebipedal gait diate between great apes and hominins
Very prognathous, receding chin, large teeth, pointed canines with gap, arcade between ape and human, hint of crest Long arms, short thumb, curved fingers and toes, bipedal
3–1.8 m.y.a. Taung Sterkfontein Makapansgat Lake Turkana (?) Omo (?) 435–530 cc; average = 450 cc Similar to A. afarensis
Less prognathous than A. afarensis; jaw more rounded; large back teeth; canines smaller than A. robustus, larger than A. afarensis; no crest —
2.5–1.6 m.y.a. Olduvai Lake Turkana Omo Sterkfontein Swartkrans 500–800 cc; average = 680 cc Limited evidence; may have been size of A. afarensis Flatter face, less sloping forehead, teeth similar to A. africanus, no crest
Limited evidence, retention of long arms, maybe retention of primitive features of hand and foot
Source: Adapted from Feder and Park, 1997.
tested. Homo erectus evolved from Homo habilis about 1.9 m.y.a., again in Africa. The earliest fossils of H. erectus are known from the eastern shore of Lake Turkana in northern Kenya. H. erectus is also the first early human form found outside Africa, in Asia and probably in Europe. In fact, a controversial date of 1.8 m.y.a. for an H. erectus fossil from China would suggest a very rapid spread of this species out of Africa. Homo erectus walked upright and had a brain size midway between that of Australopithecus and fully modern humans. The time period of H. erectus covers more than 1.5 million years, and there were a number of changes in the species during that period. In fact, there is some controversy today about the reliability of the H. erectus species designation. Remarkably, and probably relatedly, a major innovation in the tech-
nology of stone tools, the appearance of the handaxe and other bifacial tools, occurred almost simultaneously with Homo erectus in Africa (see “The Acheulean Handaxe,” Chapter 3, p. 99). The distribution of Homo erectus fossils and the archaeological evidence they produced are the subject of the next chapter. Homo erectus gradually evolved in Africa and Asia, exhibiting slowly increasing brain size, for a million years, eventually expanding into Europe. Erectus forms evolved into Homo heidelbergensis after 600,000 years ago. At some point after, perhaps, 200,000 years before the present (B.P.), Homo sapiens began to appear. Suffice it to say here that Homo erectus is the ancestor of the first Homo sapiens—and ultimately of ourselves.
There is a great deal we don’t know about our early ancestors. What would you like to know?
y
_ Concept The Family Tree
43
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 44
Site
:
Hadar A key place for finding the earliest humans All the evidence for the early hominins before 2 m.y.a. comes from Africa. Until 1970, there was relatively little evidence for the earliest human ancestors other than a few skulls and pieces of bone. The human characteristics of upright posture, large brain size, and tool use were thought to have evolved simultaneously as large primates moved out from the forest into the savanna. New fossil finds from Ethiopia, Kenya, and Tanzania have pushed back the age of the earliest known hominins and considerably
Figure 2.10 Three early hominin skulls and other remains found by Richard Leakey and his co-workers in East Africa: Homo habilis (left), Homo erectus (center), and Paranthropus robustus (right). The discovery of these species in the same geological deposits suggests that they were roughly contemporary.
_ _
44
Chapter 2 The First Humans
modified our understanding of their behavior and appearance (Figure 2.10). Shortly after 7 m.y.a., near the end of the Miocene, the first recognizable human began to appear in Africa. There are only a few examples of the earliest fossils. Most of the early hominin specimens date somewhat later, to the Pliocene and the beginning of the Pleistocene, a time often described as the Plio-Pleistocene. One of the most productive areas of research is in a region known as the Hadar (huh-DAR), northeast of the
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 45
Artifacts
city of Addis Ababa in Ethiopia (Figure 2.11). In this geologically active zone, the combination of faulting, rapid deposition, and continued erosion has exposed numerous layers from the Pliocene that contain some of the earliest human fossils yet discovered (Figure 2.12). Donald Johanson, a paleoanthropologist now at the Institute for Human Origins in Tempe, Arizona (Figure 2.13), and French geologist Maurice Taieb began a search for early hominin fossils in this area in 1972. One of the most complete early human skeletons ever discovered was found by this team in 1974. Johanson spotted a small arm bone on the ground while on a survey walk. He picked it up and immediately noticed bits of skull and bone. As he continued, he collected more pieces— fragments of vertebrae, limbs, and jaw. The next day he returned with a large crew and carefully sifted the earth at the site to recover all the fragments they could. The finds of Pliocene hominins usually consist of only a single tooth or at most a few bones. But after 2 weeks of searching and sifting,
Ash layer 2.8 ± 0.04 (K/Ar)
Lucy
Figure 2.11 The Pliocene stratigraphy at Hadar, showing the approximate age of the layers and where Lucy and the “first family” were found. These layers are exposed in the sides of gullies and on the surface by erosion. The fossils appear on the eroded surface. The number in the illustration followed by (K/Ar) provides the radiopotassium date for the layer in millions of years before the present (see “Dating Methods,” p. 48).
First family
Basalt 3.75 ± 0.1 (K/Ar)
Hadar pigs
Basal ash
Figure 2.12 The severely eroded badlands of Pliocene and Pleistocene deposits in the Hadar region of Ethiopia. This is a part of the world where many fossil hominins have been found.
_ Site Hadar
45
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 46
dating is explained later in the chapter in “Dating Methods,” p. 48.) Earlier still from Hadar, possibly dating to 3.2 m.y.a., are the skeletal parts of what has been described as “the first family”—over 200 bones from at least seven individuals, five adults and two children, found together. This is an absolutely remarkable discovery. Johanson’s own description of the discoveries conveys the excitement and the significance:
Figure 2.13 Donald Johanson (left) and co-workers at the Institute for Human Origins with some of the fossil remains from Hadar.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the Lucy specimen, see the Internet exercises on your online learning center.
_ _
polygynous Having more than
one mate.
46
Chapter 2 The First Humans
Johanson’s team found almost 100 fragments—about 40% of a complete hominin skeleton. Unfortunately, most of the skull was missing (Figure 2.14). For the first time, a reasonably intact early hominin was available, enough to reconstruct virtually the entire skeleton. Lucy, as this individual was dubbed by her discoverers, provides much new detail on the anatomy of torso and limbs. Lucy was probably female, small in stature—1.2 m (just under 4 ft) tall—small-brained, and about 20 years old when she died. Absolute age for the skeleton is estimated as somewhat more than 2.9 million years. (Absolute
Anthropology student John Kolar spotted an arm-bone fragment. From some distance away, Mike Bush, a medical student, shouted that he had found something just breaking the ground surface. It was the very first day on survey for Mike. “Hominin teeth?” he asked, when we ran to him. There was no doubt. . . . Michèle Cavillon of our motion picture crew called to me to look at some bones higher up the hill. Two bone fragments lay side by side—one a partial femur and the other a fragmentary heel bone. Both were hominin. Carefully, we started scouring the hillside. Two more leg bones—fibulae— showed up, but each from the same side. The same side? That could only indicate two individuals. . . . Time was of the essence. Rainstorms during the months of our absence could wash away fragments that would be lost forever down the ravines. . . . Each day produced more remains. . . . So we had evidence of young adults, old adults, and children—an entire assemblage of early hominins. All of them at one place. Nothing like this had ever been found! (1976, pp. 805–811) These individuals apparently died as a group, possibly in a local flash flood, and their bodies were washed to a location where their bones were preserved until the present. The sediments in which the fossils were discovered are lakeshore and riverine deposits. There are very few remains from other species in these layers and no sign that the hominin bones had been gnawed or eaten by other animals. The discovery of these bones from a group of individuals provides answers to
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 47
a number of questions. The presence of so many individuals together indicates that our earliest ancestors did indeed live in groups. The individuals were relatively small (1.2–1.5 m, or 4–5 ft, tall) but strong and sturdy. The small heads had human molars and small canine teeth. However, an apelike gap exists between the incisor teeth and the canines. The skeleton below the neck was almost completely human, but the arms were long and the bones of the hand were heavy, with large muscle attachments. The anatomy of the leg and pelvis indicates that they walked upright. Other information on early hominin patterns of growth, differences between children and adults and between males and females, the dexterity of the hands, and more, will emerge as the study of these extraordinary remains continues. In 1994, fieldworkers in the Hadar region discovered the first intact skull of an Australopithecus afarensis. This skull confirmed the upright posture of a relative of Lucy’s, dating to 3 m.y.a. More significant, the skull was a male’s and documented significant sexual dimorphism, or size differences, among these early hominins. Pronounced differences in body size between the sexes in primates is generally correlated with polygynous mating systems, in which males compete for females. Males would have been approximately 1.5 m (5 ft) tall, weighing 50 kg (110 lb); females like Lucy were 1.0–1.2 m (3.5– 4.0 ft) tall and weighed closer to 35 kg (75 lb). This difference between the sexes is also found in modern chimpanzees but is greatly reduced in more recent human species.
Figure 2.14 The skeleton from Hadar known as Lucy, one of the most complete of an early hominin. Almost 40% of the bones are present. A member of Australopithecus afarensis, Lucy is approximately 1.2 m tall and 3 million years old.
_ Site Hadar
47
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 48
Concept
<
Dating Methods Measuring the age of archaeological remains
Figure 2.15 Garniss Curtis, developer of the radiopotassium dating method, and his associate, Robert Drake (at left), examining the laboratory apparatus as a sample is heated to extract gases for the measurement of isotopes in a potassium-argon dating laboratory.
relative dating A technique used to
estimate the antiquity of archaeological materials, generally based on association with materials of known age or simply to say that one item is younger or older than another. association The relationship between
_ _
items in an archaeological site. Items in association are found close together and/or in the same layer or deposit.
48
Chapter 2 The First Humans
In the first half of the twentieth century, there was almost no way to determine the age of the remains of early humans and their artifacts. A few relative dating techniques were used to estimate the antiquity of bones and stones in a long sequence, but it was generally impossible to establish the absolute age of a layer or an artifact in calendar years. Relative dating methods often use stratigraphic relationships to sequence older and younger materials; lower layers of sediments in the ground, in caves, and elsewhere are, not surprisingly, older than the layers on top. Relative dating methods also rely on the association of various items. For example, glass bottles have changed over the past 200 years. Bottles have become seamless, with shorter necks and narrower bodies. The known dates for certain types of bottles can be used to determine the age of other materials found with them. Similarly, the thickness of the stem and bore of clay smoking pipes from the eighteenth and nineteenth centuries steadily declined as the technology for manufacturing these items improved. Thus, the age of a historical trading post in Canada can be estimated by measuring the broken pieces of pipe stem found there. A photograph can be approximately dated by the known age of an automobile or other objects in the picture. The bones of extinct animals, such as elephants in France, found with stone tools were clear evidence that the artifacts were as old as the elephants, even if the exact age was uncertain. Relative dating methods, however, are limited. It is often necessary to know
the ages of archaeological materials more precisely in calendar years— absolute dating—to answer most questions about the past. Although many methods for determining absolute dates are now available, the most common techniques rely on the properties of radioactive decay in certain elements (Table 2.2). Many elements have both stable and radioactive atomic forms, known as isotopes. The techniques used for determining absolute dates with these elements are referred to as isotopic techniques. Perhaps best known among these is radiocarbon dating, but certain other elements—potassium, uranium, and calcium, for example— can also be used for dating purposes. Radiopotassium dating, or potassium-argon dating, used to determine the age of early human ancestors and their remains, is described in the following section. Radiocarbon dating, used to date archaeological materials and events that occurred within the past 40,000 years, is discussed in Chapter 4.
RADIOPOTASSIUM DATING Radiopotassium dating, also known as potassium-argon dating, is a technique of crucial importance for determining the age of the earliest human remains (Figure 2.15). This technique can date most of the earth’s history and has been used to measure the age of the oldest rocks on our planet, as well as samples of moon rocks. The first potassium-argon dates from the lava at the base of Olduvai Gorge— 1.75 m.y.a.—startled the scientific community in the 1960s. These fossil remains were almost 1 million years older than previously believed. Other early remains in East Africa from Laetoli, Koobi Fora, Hadar, and elsewhere have also been dated using the radiopotassium technique. Bones and artifacts are not themselves directly dated; rather, newly formed volcanic rocks or ash deposits that lie directly under or over the prehistoric
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 49
TABLE 2.2 The Major Dating Methods Used in Archaeology Method
Materials
Range
Principle
Limitations
Radiocarbon
Wood, charcoal, bone, carbonate Volcanic rocks or minerals Coral, mollusks, travertine Undisturbed sediment or volcanic rocks Intact hearths, kilns, burned areas
100–40,000 years
Radioactive decay
Unlimited but approximate 30,000–300,000 years Unlimited but approximate
Radioactive decay
Contamination, calibration Appropriate samples are rare Few labs, technical problems, contamination Few labs
Radiopotassium Uranium series Geomagnetism
Archaeomagnetism
2000 years
Thermoluminescence (TL)
Pottery, heated stones, calcite
1,000,000 years
Electron spin resonance
Heated crystalline stones, calcites, bones, shell Obsidian artifacts
1,000,000 years
Tree rings in preserved logs and lumber Volcanic rocks, crystalline materials
8000 years
Obsidian hydration
Dendrochronology
Fission track
materials are analyzed. These dates thus bracket the archaeological materials in time. The technique is based on the following principles. Potassium (chemical symbol K) is found in abundance in granites, clays, and basalts in the minerals of the earth’s crust. Potassium occurs in several stable forms and has one radioactive isotope, 40K, with a half-life of approximately 1.3 billion years. Half-life is a measure of the rate of decay in radioactive materials; essentially, half of the radioactive material will disappear within the period of one half-life. Because this potassium isotope has such a very long half-life, it is usually not possible to date materials that are younger than about 500,000 years old, because too little decay would have taken place to measure. The radioactive isotope 40K decays into argon (40Ar), an inert gas, and calcium (40Ca). The materials generally dated by the 40K/40Ar technique are limited to rocks, volcanic ashes, and other substances that contain radioactive potassium and trap the argon gas that is produced. The molten state of the rock
35,000 years
100,000–1,000,000 years
Radioactive decay Alignment of particles with pole reversals Alignment with changes in location of the earth’s magnetic pole Accumulation of TL in crystals Accumulation of unpaired electrons in crystals Accumulation of weathering rind on artifact Counting of annual growth rings Radioactive decay leaves microscopic track in crystals at known rate
permits the release of trapped gas in the parent rock and resets the argon reservoirs in the new rock to zero. The 40Ar begins to accumulate as soon as a rock is formed. Using sophisticated counters that measure and record the amount of 40Ar compared with the amount of 40K remaining, researchers can determine how much 40K has decayed and thus the amount of elapsed time since the rock was created. Advances in radiopotassium dating in recent years have greatly improved the method. Today a technique called argon-argon dating is used to measure the proportion of 40Ar to 39Ar. Using this method, the stable isotope of potassium 39K is converted to 39Ar by neutron bombardment in a nuclear reactor. Both argon isotopes can then be measured from the same sample. In addition, much smaller samples can be used. The improved precision of this technique allows samples younger than 100,000 years to be dated. In one recent study, an argon-argon date of A.D. 73 was obtained for the eruption of Mount Vesuvius, which buried the Roman town of Pompeii in A.D. 79.
Few labs, calibration
Environmental irradiation rate, few labs Few labs, experimental technique Requires local calibration Region-specific
Materials rare in archaeological context
absolute dating A method of assigning archaeological dates in calendar years so that an age in actual number of years is known or can be estimated. isotope One of several atomic states
of an element; for example, carbon occurs as 12C, 13C, and 14C, also known as carbon-14 or radiocarbon. isotopic technique A method for absolute dating that relies on known rates of decay in radioactive isotopes, especially carbon, potassium, and uranium. radiopotassium dating An absolute dating technique based on the principle of decay of the radioactive isotope of potassium, 40K. Also called potassium-argon dating. half-life A measure of the rate of
decay in radioactive materials; half the radioactive material will disappear within the period of one half-life. argon-argon dating A more accurate method of potassium-argon dating that involves converting potassium to argon before the isotope ratios are measured.
Concept Dating Methods
49
_
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 50
Site
:
Laetoli Conclusive evidence of our first steps
The study of fossil origin and evolution is a kind of mystery play, acted out in remote antiquity, in many scenes and in many places. The players have long ago departed the stage, and left their all too sparsely distributed and brittle bones buried deep in the rocks in a cave. The walking shadows, the poor players who have strutted and fretted their brief hours on the stage, have in most cases been heard from no more. —Ashley Montague (1964)
Figure 2.16 Mary Leakey recording the 3.6-million-year-old footprints of Laetoli.
_ _
50
Chapter 2 The First Humans
As we have noted, the three distinctive characteristics of being human are upright posture, a large brain, and tool use. The question of which came first has been dramatically answered by a discovery in East Africa. The evidence for this new posture comes not only from the fossil bones, however, but also from actual footprints preserved at the site of Laetoli (lay-TOE-lee) in Tanzania, discovered by Mary Leakey in 1976 (Figure 2.16). Laetoli is located about 70 km (40 mi) southeast of Olduvai Gorge. Sometime around 3.6 m.y.a., an active volcano near Laetoli covered the area with a layer of volcanic ash. Following a light rain shower, various animals moved across the damp layer of ash. A chemical reaction between rainwater and the ash quickly hardened their tracks; even the impressions of the raindrops are preserved in some areas at the site. Hares, birds, extinct elephants, pigs, buffalo, rhinos, a
saber-toothed tiger, and many baboons left their footprints. The numerous sets of tracks do not often overlap one another, suggesting that this layer of footprints was quickly buried by more ash, ensuring its preservation. Radiopotassium dating determined that the age of the ash layers, and therefore the footprints, was between 3.8 and 3.5 million years (see “Dating Methods,” p. 48). Early hominins walked across the fresh ash as well (Figure 2.17). The 70 or so human footprints continued over a distance of more than 6 m (20 ft) and were made by three individuals. The longest track contains about 30 prints of an individual walking on two feet with a stride and balance that is clearly human. A second, smaller individual followed in the footprints of the first, and a third set of prints lies alongside the first. The footprints look human, with a well-defined arch and an absence
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 51
Figure 2.17 An artist’s reconstruction of the early hominins at Laetoli, walking across the volcanic ash 3.6 m.y.a.
of the diverging toe that is characteristic of great apes. Studies of the size and depth of the prints suggest that two of the individuals were approximately 1.4 m (4 ft, 8 in) tall, and the third 1.2 m (4 ft) tall. The footpath of the second individual indicates that this early hominin stopped briefly and turned slightly to the left before continuing. Mary Leakey, the excavator of these fossil footprints, speculated about the scene:
Figure 2.18 The footprints at Laetoli are clearly from bipedal individuals. The weight distribution in ape feet is along the side of the foot (left). The big toe on apes is for grasping and does not carry weight. In humans, the weight is carried from the heel, along the side, and across the ball of the foot to the big toe (right).
This motion—the pause, the glance to the left—seems so intensely human, it transcends time. Three million six hundred thousand years ago, a remote ancestor— just as you or I—experienced a moment of doubt. (quoted in Lewin, 1988, p. 57) The brain of these earliest hominins from the Pliocene was no larger than that of modern apes, nor had their teeth changed a great deal from those of their ape ancestors. Fossil remains, particularly fragments of skulls from Laetoli and elsewhere in East Africa, demonstrate that the human brain had not yet begun its major expansion. No stone tools have been found in deposits of this age at Laetoli; such equipment was apparently not yet part of the human repertoire. What was different, however, was a shift to a new form of movement.
The earliest human fossils give evidence of bipedalism—they walked on two feet, with a stride very similar to our modern one (Figure 2.18). In fact, these earliest humans might best be portrayed with the head and face of an apelike creature atop a small, upright human body, stepping into the future. As Mary Leakey went on to say, “The outstanding evolutionary question now is: What was the selection pressure that produced bipedalism?” (quoted in Lewin, 1988, p. 57).
They are the most remarkable find I have made in my entire career. . . . When we first came across the hominid prints I must admit that I was skeptical, but then it became clear that it could be nothing else. They are the earliest prints of man’s ancestors, and they show us that hominids . . . walked upright with a free-striding gait, just as we do today. —Mary Leakey (quoted in R. Leakey, 1981)
bipedalism The human method of
locomotion, walking on two legs.
Site Laetoli
_
51
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 52
Site
:
Swartkrans South African caves with many early human fossils
Figure 2.19 C. K. Brain and some of the fossil remains at Swartkrans cave. The canine teeth of the leopard jaw precisely fit the two indentations in the skull, supporting Brain’s theory that many of the early human bones are the remains of carnivore meals.
breccia The accumulated materials
_ _
from cave deposits that harden into a conglomerate rock, including sediments, rocks, and animal bones.
52
Chapter 2 The First Humans
Swartkrans (SWORT-kranz) is the name of one of several caves in the Transvaal region of South Africa that contain both human and animal fossils from the early Pleistocene. The Transvaal is a large area in the center of South Africa underlain with limestone bedrock. This soft, permeable stone is easily dissolved by running water, which creates underground rivers, caves, and sinkholes. Streams and
heavy rains from the surface wash in a variety of surface sediments, debris, bones, and other objects, which gradually accumulate in these chambers. Other agents also add materials to these deposits. African porcupines, for example, collect bones at their nests and shelters, often within the confines of caves; hyenas carry bones and other objects to their dens. Some birds, such as owls and other birds of prey, regurgitate small pellets containing the bones of the small rodents they consume. The accumulated materials eventually harden into a fresh rock known as a conglomerate, or breccia. Early fossils were first discovered when the breccias in these caves were commercially quarried for lime, used for making plaster and cement. Dense bone concentrations were discarded to obtain purer deposits. Eventually, the antiquity and the importance of these bones were recognized, and expeditions from museums and universities in South Africa and elsewhere conducted excavations to uncover more fossil materials. Collectively, these excavations have uncovered the remains of more than 150 individual hominins of several species. There are four major problems associated with studying these materials: (1) extracting the fossils from the rock in which they are encased, (2) estimating the age of the fossils and the deposits, (3) determining the cause of death of the animals and how they got into the caves, and (4) designating the genus and species for the remains. Removal of the fossils from these deposits is very difficult. The fossil bones are embedded in breccia and in fact have become rock themselves. Excavators have used jackhammers and even dynamite to break apart the breccia to get at the mineralized bones.
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 53
In recent years, large blocks of the breccia have been taken to a laboratory where acid treatments gradually dissolve the rock, thereby allowing the removal of the fossils. Because these materials are buried in the earth, geological forces have modified and distorted the stone and the fossils. The weight of deposits sometimes flattens or warps bone or crushes materials together. The limestone deposits in these caves provide little that can be directly dated. There is no fresh volcanic ash or rock for potassium-argon dating. Most of the dates from the caves are based on the known ages of extinct animals that once existed throughout much of Africa. For example, changes over time in the size of the teeth of fossil pigs can be used to estimate the ages of the various deposits. Such dating is done by the association of remains with pig teeth of known age found in the same breccia. On this basis, the early Pleistocene breccia at Swartkrans was thought to date to between 1.7 and 1 m.y.a. However, new dating methods suggest that the earliest deposits may be as old as 3 million years, to between 2 and 4 million years, making them comparable in age to many of the finds from East Africa. C. K. Brain, a paleontologist in South Africa, has been excavating at the site of Swartkrans for more than 20 years (Figure 2.19). Brain notes two major episodes of breccia accumulation, apparently of short duration, on the order of 10,000 years each. Dating of the deposits is complicated by the manner in which the breccias form. There is no simple stratigraphy in which materials are piled up in regular horizontal layers; continuing water flow erodes and partially dissolves older deposits, then new episodes of accumulation fill the cavities in the old deposits. Each cave thus contains a honeycomb of deposits from various periods. These deposits are the source for fossils of both Paranthropus robustus (a late australopithecine) and Homo erectus. The partial remains of at least 80 P. robustus individuals have been re-
covered from the deposits at Swartkrans, along with fragments of bone and teeth from 6 individuals of H. erectus. These two hominin forms appear to be contemporaneous at Swartkrans, as at Olduvai. This astounding number of individuals from a single deposit raises the question of how so many came to be in one place—a question that has yet to be fully answered. The shape and structure of the hand bones from several P. robustus individuals suggest that this hominin made and used some of the stone tools that are also commonly found in the cave deposits. In addition to the stone tools, there are bone artifacts with polishes that resulted from their being used as digging implements, probably to obtain the roots and bulbs of plants growing in the area. Recent data from Swartkrans, from approximately 1 m.y.a., suggest that the individuals in the cave used fire, perhaps to keep preying leopards at bay. If this evidence is reliable, it records the earliest instance of the intentional use of fire. There is also some evidence for the use of fire at Koobi Fora in East Africa perhaps 1.5 m.y.a., but this is less certain. How the bones of so many early hominins ended up in the deposits also is not clear. This question has been the subject of lengthy debate. Raymond Dart, an anatomist in South Africa, studied the human and animal remains from the cave breccias for almost 50 years until his death in 1988. He proposed a number of theories about the life and times of Australopithecus. Professor Dart, the original discoverer of the australopithecines, believed that the early hominins were extraordinary predators, hunting many animals and bringing their carcasses back to the limestone caves. Dart argued that animal bones found in the caves were the remains of meals and evidence of the prowess of Australopithecus as a hunter of even the largest animals. Dart’s graphic depiction conveys his image of our earliest ancestors: “Man’s predecessors . . . seized living quarries by violence, battered them to death, tore apart their broken bones, dismembered them
Figure 2.20 An artist’s interpretation of the Swartkrans leopard hypothesis. A Pliocene leopard consumes a hapless australopithecine in a tree above the entrance to the Swartkrans depression. (From C. K. Brain, The Hunters or the Hunted? Copyright © 1981 University of Chicago Press. Used with permission of the publisher, University of Chicago Press.)
Why was Africa such a propitious place for the emergence of humanity?
y
_ Site Swartkrans
53
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 54
limb from limb, slaking their ravenous thirst with the hot blood of victims and greedily devouring living writhing flesh” (1953, p. 209). Dart further suggested that the high number of skulls in the caves, often with the bases broken away, was evidence that early humans also hunted their neighbors. The question is one of hunter or hunted, according to Brain. Brain’s detailed studies provide numerous insights about the processes of deposition in the limestone caves, and they contradict some of the arguments of Dart. Many of the cracks and breaks in the fossil remains that Dart interpreted as evidence of violent death more likely resulted from geological distortion after burial, according to Brain. Brain argues that the vast majority of larger fossils came into the cave as the remains of carnivore meals, especially of leopards. Brain points out that leopards often drag their prey into trees, out of reach of other predators
_ _
54
Chapter 2 The First Humans
(Figure 2.20). The damp entrance shafts to caves such as Swartkrans would likely have been overgrown with trees, which may have been lairs for satiated leopards. Over 40% of the human fossils are immature individuals who may have been vulnerable to predator attacks. Brain also points to a skull with two crushing indentations in the forehead, punctures that fit precisely with the huge canine teeth of a leopard. Investigations at this important South African cave document the contemporaneity of Paranthropus robustus and Homo erectus. Swartkrans finds demonstrate that P. robustus almost certainly made and used stone tools, that bone tools were an important part of the equipment for these early hominins, that plants were probably important in the diet, and that fire may have been used, at least for defensive purposes.
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 55
Concept
<
Hunters or Scavengers? Ways our early ancestors obtained their food
What the early human diet consisted of is difficult to determine, because the remains of meals are generally not well preserved. The wild plants of East Africa probably provided a ready source of food for early humans, but plant materials are not preserved. The relative importance of fruits, nuts, and other plant foods in the diet is not yet known. Another difficult matter, and a major controversy in paleoanthropology, is the issue of scavenging versus hunting: How did early hominins obtain meat? Some scholars believe that the first humans were primarily scavengers, visiting the kills of lions and other predators, and competing with hyenas and vultures for the morsels that remained. These scholars argue that the actual hunting of large animals is a relatively recent development in human prehistory. Others contend that early hominins were in fact hunters—stalking, killing, butchering, and eating the creatures of Plio-Pleistocene East Africa. The evidence is scanty and open to debate. Only a few facts are known: Chimpanzees and baboons occasionally hunt, kill, and consume small animals, and, by the end of the Pleistocene, humans were major predators and large-game hunters. Evidence from numerous PlioPleistocene sites certainly suggests that humans brought various animal parts back to a common location and removed the meat and marrow with stone tools. In one instance, the large leg bone of an antelope-size creature was broken into ten pieces, in a fashion that modern hunters in the area today use to obtain the nutritious and tasty marrow. On this same bone, tiny scratches made by stone tools are also visible. Such cutmarks are sometimes found at places where large pieces of meat have been removed from the bone, suggesting that the animal was butchered for the meat (Figure 2.21). This evidence may indi-
cate that early hominins were hunters with access to the best cuts of meat from their prey. But paleoanthropologists Richard Potts and Pat Shipman have observed that cutmarks are sometimes found over the marks left by the teeth of carnivores, suggesting the scavenging of animal carcasses. It seems clear that stone tools were used to butcher meat, but neither the presence of cutmarks indicating meat removal nor evidence of the extraction of marrow from bone demonstrates that the food was hunted, not scavenged. Recent studies of marks on bones from Olduvai Gorge suggest a sequence of large carnivore teeth, cut by stone tools, overlain by smaller carnivore tooth marks. Such a pattern suggests that the animal was killed by a large carnivore, scavenged by humans, and then eaten by smaller animals. Comparisons of the kinds of animals represented at Olduvai Gorge and elsewhere with the prey of modern hunter-gatherers in the Kalahari Desert of South Africa support the argument that australopithecines were hunters. The range of prey types and their sizes are very similar in the two locations, suggesting that the Olduvai hominins may have been hunters, like people of the Kalahari, rather than scavengers. Still, none of these pieces of evidence is definitive. Evidence for sites where large game animals were killed is not incontrovertible until much later in the Pleistocene, after 500,000 years ago.
(a)
(b)
Figure 2.21 Cutmarks on bones from Koobi Fora are a strong argument that our early ancestors were meat-eaters and hunters as well as scavengers. (a) A photomicrograph of round-bottomed grooves made by hyena teeth on modern bone. (b) A photomicrograph of a V-shaped cutmark made by a stone tool on modern bone.
cutmark A trace left on bone by a
stone or metal tool used in butchering a carcass.
Concept Hunters or Scavengers?
55
_
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 56
Site
:
Olduvai A trail of biological and behavioral evolution from the early Pleistocene to the recent past
Figure 2.22 Olduvai Gorge, cutting 100 m into the Serengeti Plain and 2 million years into human evolution.
_ _
56
Chapter 2 The First Humans
Flying low across northern Tanzania, one crosses an enormous wilderness of grassland and solitary trees, a region filled with herds of wildebeest, giraffes, elephants, and many other animals. This is the fabled Serengeti (ser-in-GETee) Plain—the place of safari. The level surface of the plain results from the long, gradual accumulation of geological sediments, especially volcanic materials such as ash and lava. Two million years ago, this area was a large
bowl-shaped basin, ringed by a series of volcanic mountains and uplands. Active volcanoes filled the air with ash and covered the ground with molten lava, which hardened into new rock. The basin trapped rainfall, forming lakes and wetlands during the beginning of the Pleistocene. Silts and sands, carried by running water, were deposited in these lakes, which grew or disappeared over time as rainfall amounts varied with changes in cli-
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 57
mate. Along the shores of these lakes, creatures of the early Pleistocene in East Africa found food, reproduced, and died; occasionally, their bones were buried and preserved in the accumulating layers of sediment. The richness of the lakeshore environment is represented by the abundance of fossil animal bones that are found there. Antelope, giant buffalo, and wild sheep occur in large numbers, along with aquatic animals, such as the giant crocodile, the hippopotamus, and various species of fish and fowl. The layers of lava, ash, and lake deposits continued to build up until the basin became relatively level, resulting in the surface of the Serengeti Plain today. About 200,000 years ago, a particularly violent series of earthquakes and volcanic activity opened a crack in the surface. Seasonal streams cut and eroded a large gully into the layers of sediment. Gradually, a canyon, some
40 km (25 mi) long and almost 100 m (325 ft) deep, wound its way from the top of the Serengeti Plain through the layer cake of deposits. This canyon is Olduvai (ol-dew-VIE) Gorge, one of the most famous prehistoric sites in the world (Figure 2.22). Each step down into the gorge takes us back 6000 years in time, toward the layer of basalt at the very bottom, dating to 1.9 m.y.a. (Figure 2.23). Along the steep sides of this gorge, two archaeologists—Louis and Mary Leakey—began an extended vigil, in quest of the remains of the earliest humans. Starting in 1931, Louis and, later, Mary Leakey made the arduous journey from Nairobi each summer to spend several weeks at the rugged exposures of Olduvai. Accompanied by their dogs, and later their several children, they searched for fossil hominins. Louis Leakey had found numerous crude stone tools in the
100 m
Years B.P. 400,000
Handaxes Homo erectus Bed IV
72 m
700,000
60 m
1,200,000
32 m
1,650,000
Figure 2.23 A schematic cross section through the 100 m of deposits at Olduvai Gorge, naming the various fossil forms and types of stone tools, with approximate ages.
Bed III Handaxes, pebble tools Homo erectus Homo habilis Bed II Pebble tools
Homo habilis Paranthropus boisei 1,800,000
Bed I 0m
Basal basalt layer
1,900,000
_ Site Olduvai
57
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 58
Figure 2.24 Some of the more common early Pleistocene animal species at Olduvai Gorge.
Okapi
Oryx
Waterbuck
Porcupine
Sivatherium
lower layers at the gorge and was convinced that the bones of the toolmakers would also appear in this remarkable series of deposits. Not until 1959, however, 28 years after Louis’ first visit, was the persistence of the Leakeys rewarded by Mary’s discovery of a very early fossil, initially named Zinjanthropus. At the time, the fossil was thought to be approximately 1 million years old—twice the age of the then-earliestknown remains from Java. Zinj, as this fossil is affectionately known (or Paranthropus boisei, as it is scientifically termed), actually dates closer to 2 million years old. The Leakeys’ discovery brought the search for the first humans to Africa and eventually back into the Pliocene epoch. Their discovery of Zinj also brought world recognition for their efforts in the form of acclaim and funding, which supported more extensive investiga-
_ _
58
Chapter 2 The First Humans
Deinotherium
tions at Olduvai. With that funding, the Leakeys were able to examine a larger area of the gorge in 2 years than had been possible in the previous 20 years. That intensive work paid off in the discovery of more fossils and a whole series of archaeological sites. Very old standardized objects of human manufacture (stone artifacts) appeared in the lower layers at Olduvai. Olduvai provided the first clear documentation that crude stone tools and the bones of very early hominins occurred at the same point in geological time. Over 70 prehistoric localities with stones or bones, or both, have been recorded in the geological layers of the gorge to date; perhaps 10 of these represent actual living areas where tools were made and used. Some of the stones are unmodified and may have been used as anvils and for other purposes. Other stones were in-
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 59
Figure 2.25 Typical Oldowan pebble tools, shown with a tennis ball for scale.
0
tentionally bashed with another stone to shape and manufacture tools (Figure 2.25). These stone artifacts had strong, sharp edges, providing cutting equipment for a species lacking sharp teeth or claws. The materials for these artifacts were often brought from the rocky hills some 10 km (6 mi) away. Raw materials were selected on the basis of specific properties. Fine-grained stone was used to make small cutting tools, and basalt and quartz were used for heavy chopping equipment. Tools described as choppers, spheroids, and discoids were created by knocking off flakes of stone from a rounded cobble or large pebble. These sharp-edged cobbles are about the size of a tennis ball and are known as Oldowan pebble tools, named after the gorge itself. The flakes that had been struck off these pebble tools also had sharp edges and were likely used as tools. One of the Olduvai sites contains a large quantity of broken and fragmented bone, along with stone tools. Many of the bone fragments are clustered in an area of about 5 × 10 m (16 × 33 ft, the size of a large room), with an empty zone several feet wide surrounding this concentration. Perhaps a thorn hedge or barricade was placed in this area to protect the inhabitants in the center.
5 cm
At another site in Olduvai Gorge is a group of several hundred rocks in a roughly circular arrangement, surrounded by the bones of giraffes, hippopotamuses, antelopes, and elephants (Figure 2.26). The reason for such concentrations is unknown; it is not even clear whether early hominins were responsible for killing the animals represented by the bones. However, the hominins almost certainly collected the bones. Two other sites at Olduvai are known to have been places of animal butchering. At one of the sites, known as FLK North, the bones of an elephant lie scattered on the ground along with stone artifacts. The elephant would have been much too heavy to move and was very likely butchered at the spot where it died. Most of the bones from the elephant are present, disarranged by the butchering and surrounded by stone tools and flakes. Striations and cutmarks on the bones document the use of stone flakes to remove meat from the skeleton. (The issue of whether the hominins at Olduvai actually hunted these animals was discussed earlier; see “Hunters or Scavengers?” on p. 55.) Other evidence suggests that most of the living floors at Olduvai were occupied during the wet season. Tortoises hibernate during the dry season, making them difficult to capture, yet their remains are common at most of
Oldowan The name given to the assemblages of early pebble tools and flakes belonging to the Basal Paleolithic, derived from Olduvai.
_ Site Olduvai
59
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 60
Figure 2.26 The plan of part of an excavated deposit at Olduvai, containing a concentration of elephant and other bones, with stone tools shown in solid black. This site likely represents the place where parts of these animals were butchered.
D
D Fem
ur
Pelvic girdle
Pelvic girdle A
Femur
Uln
Hum
a
erus
Ribs
Hu
me
ru
s
Scapula a
ul
ap
D
Sc
Mandible 0
1m
Figure 2.27 The Leakey family at work in Olduvai Gorge, ca. 1960.
the sites at Olduvai. Such information suggests that our early ancestors may have been absent from the Olduvai lakeshore during the dry season, pursuing other activities and perhaps game elsewhere in the region. Olduvai will remain one of the most important archaeological sites in the world because it contains the in-
_ _
60
Chapter 2 The First Humans
formation that helps answer many questions—the human fossils, the early Pleistocene deposits, the association of human bone and stone artifacts, and the fact that these materials are sometimes found where they were dropped by our early ancestors.
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 61
Concept
<
The Leakey Family A dynasty of paleoanthropologists
Born to British missionary parents in Kenya in 1903, Louis Leakey became interested in prehistoric artifacts as a child. At age 16, he left for England to attend Cambridge University. A rugby injury forced a break in his studies, and he returned to East Africa on a year-long fossil-hunting expedition, engendering his interest in bones and paleontology. Leakey then returned to Cambridge, taking high honors in both archaeology and modern languages. Leakey then received a 6-year research fellowship from Cambridge to conduct archaeological investigations in East Africa. One of his first expeditions centered on Gamble’s Cave and provided the first sequences for the development of stone artifacts in East Africa. In 1931, Leakey led his third expedition to Olduvai Gorge. Here, early handaxes were found soon after his arrival, and a fragment of a human jaw was discovered nearby. Leakey argued for a great antiquity for this bone, but a British geologist, Percy Boswell, accused Leakey of incompetence in a letter to a scientific magazine. This incident left Leakey with a determination to be more exact in his statements. Much of Leakey’s research was guided by his belief that Africa was the cradle of the human race, in contrast to the generally accepted scientific view that East Asia held that distinction. One of the members of Leakey’s fourth expedition was Mary Nicol, an archaeologist and illustrator, who would become his second wife in 1936. Together, Mary and Louis Leakey began their lifelong research on early humans in East Africa (Figure 2.28). The Leakeys continued their investigations at Olduvai Gorge because of its abundance of stone tools. Leakey was convinced that human fossils would also be found there, demonstrating the fact that early hominins had made the stone tools.
Leakey was appointed director of the National Museum of Kenya in Nairobi in 1945, and each summer the Leakeys would return to search for evidence of early humans in the layers of Olduvai Gorge. Finally, in 1959, after more than two decades of dogged and difficult prospecting, Mary Leakey spotted the fragments of the heavy jaw
Figure 2.28 Louis Leakey and the remains of Homo habilis from Olduvai Gorge. This discovery demonstrated that Africa was the cradle of early human evolution and brought fame to the Leakeys.
_ Concept The Leakey Family
61
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 62
of Zinj, dating to the beginning of the Pleistocene. The Leakeys proved that East Africa had indeed witnessed the emergence of the human species. In addition to the many important discoveries they have made, the Leakeys pioneered the study of early stone tools and animal remains in East Africa. Louis Leakey died in London on the first of October 1972. After Louis’ death, Mary Leakey and her children continued the search for human origins in Africa. Mary excavated at Olduvai, Laetoli, and elsewhere in East Africa, greatly enriching the record of early fossil animals and humans in this area. She compiled most of the important published descriptions of the prehistory of Olduvai. Mary Leakey died in 1996 at the age of 83. Her son Richard became the di-
_ _
62
Chapter 2 The First Humans
rector of the National Museum of Kenya in Nairobi and a very famous paleoanthropologist in his own right. In spite of several calamities, including the loss of both legs in an airplane crash, Richard Leakey continues as a champion of the peoples and animals of his native Kenya. His debates with other scientists over the meaning of the fossil finds are the subject of several popular books. Richard’s wife, Meave, is the head of the paleontological section of the National Museum of Kenya and has been involved in the search for fossil humans in the area around Lake Turkana for many years. Her daughter, Louise, a Ph.D. in archaeology from University College, London, is continuing the family tradition as director of the Koobi Fora project in northern Kenya.
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 63
Concept
<
The First Tools Simple, intentionally broken pebbles: the earliest preserved evidence for toolmaking
Simple stone tools are the earliest human artifacts that archaeologists study. It is very possible that the first humans, like chimpanzees today, used tools of wood or bone. As our closest relatives, chimpanzees provide a good model for early humans. Chimpanzees in the wild, in fact, have frequently been observed killing and eating other animals and making and using tools. In West Africa, chimpanzees often use stone or wooden objects as hammers and anvils to open nutshells. They use twigs to remove marrow from bone cavities or to extract termites from their nests. Similarly, unmodified wood, bone, and stone objects may have been used long before the appearance of deliberately modified stone tools, but these either cannot be identified or have not survived. Intentionally modified stone tools appeared first in Africa between 3 and 2 m.y.a., probably associated with the increasing importance of meat in the human diet. Such sharp flakes and cobbles provide access to the carcasses of animals, enabling early humans to cut through thick, tough skins to remove the meaty tissue—actions simply not possible without some kind of sharp implement. Stone tools provide useful cutting edges for a species that lacks both sharp teeth and claws for slicing meat, shredding plants, or digging. As Nicholas Toth says, “Sharpedged stones became the equivalent of canines and carnassials [meat-cutting teeth] and heavier rocks served as bone-crunching jaws” (1987, p. 121). The oldest dated stone tools were found in the 1990s along the Gona River in central Ethiopia, dating to 2.6 m.y.a. (see Figure 2.29). Cutmarks, made by stone tools, were found on animal bones at these sites, documenting their use for butchering carcasses. The
earliest stone tools are remarkably simple, almost unrecognizable unless found together in groups or next to other objects. Small round cobbles, the size of a large egg or a tennis ball, 5–8 cm (2–3 in) in diameter, weighing about 1 kg (1–2 lb), were collected from streambeds, lakeshores, and beaches. Early stone tools were created by striking one stone against another. This process, called percussion flaking, results in a flake being removed from the parent cobble, or core, by a blow from another stone, called a hammerstone, or some other hard object (Figure 2.30). Both the flake and the core then have fresh surfaces with edges sharp enough to be used for cutting. Initially, the fractured cobbles themselves were thought to be the intended artifact and were called pebble tools. The flakes were thought to be by-products of the manufacturing process, a kind of waste material, often referred to as débitage.
Figure 2.29 One of the earliest known stone tools from the Gona River region of Ethiopia, 2.6 m.y.a.
percussion flaking A technique for producing stone artifacts by striking or knapping crystalline stone with a hard or soft hammer. flake A type of stone artifact produced by removing a piece from a core through chipping. core The stone from which other pieces or flakes are removed. Core tools are shaped by the removal of flakes. hammerstone A stone used to knock flakes from cores. débitage A term referring to all the
pieces of shatter and flakes produced and not used when stone tools are made.
Concept The First Tools
63
_
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 64
Hammerstone
Figure 2.30 Percussion flaking, used by early hominins to create cutting edges on stone tools. One stone (the hammerstone) was bashed against another (the core) to remove one or more flakes.
Core
Striking platform
Striking point
Outer surface of core
Flake Flake scar
Hello to all intelligent life forms everywhere . . . and to everyone else out there, the secret is to bang the rocks together, guys. —Douglas Adams (1980)
flint A fine-grained, crystalline stone
that fractures in a regular pattern, producing sharp-edged flakes. flintknapping The process of making chipped stone artifacts; the striking of stone with a hard or soft hammer. assemblage The set of artifacts and
other remains found at an archaeological site or within a specific level of a site. lithic Pertaining to stone or rock. unifacial A term describing a flaked
stone tool in which only one face or side is retouched to make a sharp edge. bifacial A flaked stone tool in which both faces or sides are retouched to make a thinner tool. retouching The shaping or sharpening of stone artifacts through percussion or pressure flaking.
_ _
handedness Preferential use of the
right or the left hand.
64
Chapter 2 The First Humans
But it is now clear that flakes were equally important as cutting tools and tools for making other tools, such as shaping wood, bone, or antler into new forms for new purposes. The simple action of striking one stone against another to remove a flake and create a sharp edge was a very successful invention, one that was used and refined for more than 2 million years, until the introduction of metals just 6000 years ago. The best raw materials for stone tools during the early Pleistocene were brittle enough to break but hard and smooth enough to provide a cutting edge. The stone also had to be finegrained so that it would break in a predictable fashion, resulting in large flakes, rather than hundreds of shattered fragments. During this time, various rocks were used, including basalt (a hard volcanic lava), quartzite, and flint. Flint is one of the best and most common materials used for making stone tools. The term flintknapping is often used to describe the process of making stone tools. The kinds of pebble tools found at Olduvai, at Koobi Fora, and elsewhere in Africa are described as Oldowan, a tradition of toolmaking. The term Oldowan is applied to the entire group, or assemblage, of different stone objects found together at sites from the end of
the Pliocene and the first part of the Pleistocene. These stone, or lithic, artifacts include both unifacial pebble tools, flaked on one side only, and bifacial pebble tools, flaked on both sides. The flakes are also occasionally further modified by additional flaking, or retouching, along their edges, to shape them. We have learned other kinds of information from stone artifacts, the most durable of the remains of our ancestors. Studies by Toth have suggested that many of these flakes were made by holding the core stone in the hand and striking it with a hammerstone. Indeed, Toth argues, on the basis of the shape of the flakes, most of the flakes from Koobi Fora were produced by right-handed individuals. This handedness is not seen in the very earliest stone tools from 2.6 m.y.a. but emerged in the period between 1.9 and 1.4 m.y.a., probably correlated with the changes in the organization of the brain. The brain of modern humans is divided into two hemispheres that control different areas of thought and behavior. Handedness is a result of this lateralization of the brain. In right-handed people, the left hemisphere controls sequential abilities, such as speech, and
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 65
more quantitative activities, and the right hemisphere regulates spatial conceptualization and more abstract behavior. These functions are reversed in left-handed people. Thus, the predominance of right-handedness in stone tool manufacture after 2 m.y.a. suggests that this organizational change had already taken place. Studies by Nicholas Toth, of Indiana University, and Lawrence Keeley, of the University of Illinois–Chicago, have provided important clues about the use of these stone tools. Keeley used a high-powered microscope to examine the edges of stone artifacts from Koobi Fora. Experimental work demonstrated that different materials leave different kinds of traces in the form of polish on the edges of tools. At a magnification of 400×, Keeley observed microscopic polish and wear, indicating the cutting of meat, the slicing of soft plant material, and the
scraping and sawing of wood, on about 10% of the flakes. Two of the flakes with evidence of meat butchering were found within 1 m (3 ft) of a large herbivore bone exhibiting cutmarks. Such evidence strongly supports the use of these stone artifacts as butchering tools. Evidence of woodworking suggests that wooden tools were also being made, perhaps crude digging sticks for finding roots and tubers. This indirect information is the only evidence for the use or consumption of plant materials. Tools provide an interface between humans and the environment, enabling us to manipulate and change our surroundings. One of the most remarkable things about stone tools is the investment they represent—a vision of the future, an anticipation of action. An object was made at one time and place, often intended to be used later elsewhere.
tool Any equipment, weapon, or object made by humans to change their environment.
_ Concept The First Tools
65
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 66
Images and Ideas Bones, Stones, and Human Behavior The evolution of our biology and our activities
A
s we have noted, the late Miocene and Pliocene evidence for early hominins consists of pieces of bones and teeth, along with some small broken stones, from a few locations in Central Africa, East Africa, and South Africa. Although the evidence is fragmentary and rare, it is still possible to put together a picture of the evolutionary history of the hominins during the Miocene, the Pliocene, and the early Pleistocene. This period extends from approximately 7 to 2 m.y.a., from the earliest humans to the emergence of Homo erectus. Many new varieties of African apes appeared in the late Miocene as the climate of this region became drier and the tropical rain forests shrank. Certain of these apes were becoming bipedal by the end of the Miocene. Only a very few examples of these Miocene hominins are known, including Sahelanthropus tchadensis, Orrorin tugenensis, and Ardipithecus ramidus. Their known characteristics reveal a mix of ape and more human features. A variety of hominin forms appeared in the Pliocene, probably in response to the continued climatic trend to drier conditions. These new forms include the australopithecines and Kenyapithecus. Their bipedal locomotion is well documented, both in the fossil bones themselves and especially in the remarkable footprints from Laetoli, dating to 3.6 m.y.a. They were generally small-brained with large teeth. Sometime between 3 and 2 m.y.a., our ancestors began to make stone tools, and near the end of the Pliocene, human evolution took two major paths. Paranthropus followed a dead-end street and disappeared before 1 m.y.a. Homo habilis started along the pathway to the present, a curious species with a mixture of traits, combining large teeth and small brains, with other, more evolved features. Determination of the correct species for habilis—whether Australopithecus or Homo—remains difficult and debated. Whatever the case, habilis gave rise to erectus, a definite member of the genus Homo, exhibiting larger brains, thicker enamel, and a reduction in sexual
1500
Figure 2.31 Brain size and body weight in the higher primates and hominins. Body weight is estimated for the fossil forms. Notice the increase in both body weight and brain size in the line of human evolution.
Homo sapiens
Homo erectus
Brain size (cc)
1000
Homo habilis 500
Australopithecus africanus
Paranthropus robustus
Gorilla Orangutan
Chimpanzee Gibbon
25
_ _
50 Body weight (kg)
66
Chapter 2 The First Humans
75
100
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 67
dimorphism. The rise of erectus also saw a doubling in brain size, a dramatic geographic expansion of the genus, and new species or subspecies (neanderthalensis) (Figure 2.31). Homo erectus was essentially human below the neck. The final stages in the evolution from erectus to sapiens took place in the cranium. These individuals almost certainly lived in small groups, evidenced by the clusters of animal bones and stone tools at Olduvai, Koobi Fora, and other sites. Early hominins ate both meat and plants. How they obtained the meat is unclear, but it appears that they may have hunted small animals and scavenged the marrow-rich bones and remaining meat from large-animal kills of other predators. Beyond these few indications, however, there is very little direct evidence for reconstructing our early pedigree and behavior. The fact that there are any reasonable explanations of our earliest evolution is testimony to the diligence and ingenuity of the scientists who study those remains. In spite of the sparseness of evidence for the biology and behavior of early hominins, there are still a number of theories about how and why we became human. This understanding comes partly from comparisons with our nearest relatives, the chimpanzees, and partly from inferences made from the evidence of the stones and bones themselves. Much of the following discussion involves speculation about the evolution of human behavior, based on little evidence and vulnerable to considerable revision and modification. The brain size of modern humans is roughly 1000 cubic centimeters (60 cu in) larger than that of other primates (say, a large grapefruit compared with an orange) (Table 2.3). We have a much greater ratio of brain to body size than almost any other species, emphasizing the importance of our brain for survival. Brains are costly organs, requiring about 20% of the body’s energy production to operate. Brain tissue needs more than 20 times the energy of muscle tissue to function. This means that humans may need food that is denser in calories and nutrients than most other species need. Yet human infants are born with a remarkably small, underdeveloped brain, only some 25% of the adult size. Whereas a chimpanzee’s brain at birth is about 65% of fully adult size, the human brain grows and develops largely after birth, because of the narrowness of the birth canal. Having an underdeveloped brain, a human infant requires a long period of maternal care and attention. In contrast, a newborn foal, for example, can run and feed itself within a few days. In humans, the long period of infant dependency fosters a strong bond between mother and child and also permits children to know their brothers and sisters. Kinship and learning may well be enhanced as a result of such extended and intimate relationships. A lengthy period of maternal care, however, is expensive. Small children limit the mobility and activities of the mother, including the obtaining of food. Extended care and brain development also mean longer nursing and delayed weaning. In turn, such behaviors require more food for the mother; nursing alone may increase her metabolic demands by 50% or more. The average life expectancy among the australopithecines appears to have been low, around 20 years. Such an early age of death very likely meant that children were often orphaned and had to be nurtured by the group rather than by a single parent; communal childcare and feeding may well have been concomitant with the advance of our species. Differences between humans and the apes are also represented in the secondary sex characteristics that differentiate males and females. Human female breasts are significantly larger than those of the male, a difference more pronounced than in most other species. Breast size is obviously not related to the effective feeding of offspring. The human male penis is far larger than that of other primates, including gorillas. Human females have softer skin and a higher voice and lack much of the body hair that often characterizes other primate females.
TABLE 2.3 Average Cranial Capacity of Some Primates and Fossil Hominins
Species
Cubic Centimeters
Modern humans Neanderthals Homo erectus Homo habilis Australopithecines Gorillas Chimpanzees Orangutans Gibbons
1300–1400 1300–1450 800–1200 500–800 350–550 500–570 350–450 360–460 80–120
Images and Ideas Bones, Stones, and Human Behavior
_
67
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 68
© 2004 by Sidney Harris. Reprinted with permission of ScienceCartoonsPlus.com.
y
What were the conditions that gave rise to the steps toward human evolution?
With the stimulus of constantly available sex, protohominins had begun the most fundamental exchanges the human race would ever make. Males and females were learning to divide their labor, to exchange meat and vegetables, to share their daily catch. Constant sex had begun to tie them to one another and economic dependence was tightening the knot. —Helen Fisher (1982)
estrus The cycle of female sexual
receptivity in many species of animals.
_ _
68
Chapter 2 The First Humans
The vagina in the human female points sufficiently toward the front of the body to permit face-to-face sexual intercourse. Another biological difference lies in the absence of an estrus cycle in human females. The female members of most animal species are sexually active only for limited intervals, a period known as estrus, a few days each month or each year. Female apes display sexual receptivity by means of flaring or brightly colored sex organs, but ovulation is concealed in human females. Clearly, there have been strong evolutionary selective forces favoring these characteristics. Aside from biological similarities to and differences from other animals and our closest relatives among the apes, are we indeed closer to the angels? We possess other attributes that distinguish us from the animals. Our large brain and intelligence enable us to make decisions and act rationally, differently than the way our basic instincts might drive us to behave. Humans have moved from purely instinctual behavior to reason and thought. We may flee a fire, but we may also turn back into that fire to save others. Humans have technology; we make and use a wide range of tools and other devices that increase our chances for survival. Tool use by humans is often considered to be the most distinctive characteristic of our species, even though other animals may use or even make simple tools. For example, the sea otter wields a rock to break open the shell of an abalone. Anthropologist Jane Goodall has observed chimpanzees using a variety of tools: thrashing about with branches for display, using clubs and missiles for defense, and selecting a twig and stripping its bark to probe the nests of termites and attract them to the stick in order to eat them. West African chimps use stone and wooden hammers to crack and open nutshells. Although tool use may not be unique to humans, using tools to make other tools does distinguish the human animal. The evidence from Africa clearly indicates that our first step on the road to becoming human was the shift to bipedalism. It now seems clear from the African evidence that we had upright posture for several million years before either the use of stone tools or significant increases in the size of our brain. Apparently, standing upright was very important to our early survival.
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 69
Explanations for bipedalism focus on whether the feet or the hands were changing. Mary Leakey and her colleagues have suggested that we became upright to pursue migratory animal herds on the savanna. Bipedalism and powerful strides would be advantageous for moving long distances to follow herds for meat. And upright posture would free the hands for carrying young offspring. This hypothesis ignores the fact that australopithecine teeth seem adapted to an increasingly vegetarian diet, becoming bigger and flatter over time. Walking around on two feet could, however, be useful for collecting plant foods scattered across the landscape. Other ideas regarding the evolution of bipedal locomotion relate to food getting, carrying, or sharing. Clifford Jolly, of New York University, has proposed that bipedalism was adaptive because it provided a means for gathering and eating young leaves, seeds, and pods of the African thornbush growing on the African savanna. Owen Lovejoy, of Case Western Reserve University, has proposed that our ancestors became two-footed because bipedalism enabled males to carry food back to a favored location to be shared with females, a concept known as male provisioning; this would have increased the birth rate and reproductive success of humans because females could support more than one dependent offspring at a time. Lovejoy wrote: Both an advanced material culture and the Pleistocene acceleration in brain development are sequelae to an already established hominin character system, which included intensified parenting and social relationships, monogamous pair bonding, specialized sexual-reproductive behavior, and bipedality. . . . The nuclear family and human sexual behavior may have their ultimate origin long before the dawn of the Pleistocene. The proposed model accounts for the early origin of bipedality as a locomotor behavior directly enhancing reproductive fitness, not as a behavior resulting from occasional upright feeding posture. It accounts for the origin of the home base in the same fashion as it has been acquired by numerous other mammals. It accounts for the human nuclear family, for the distinctive human sexual epigamic features, and the species’ unique sexual behavior. (1981, p. 350)
We do not yet know why, in fact, certain apes became bipedal—and thus human. In the final analysis, however, it is culture that distinguishes the human creature. Culture is what anthropologist Leslie White called our “extrasomatic means of survival”—the nonbiological, nongenetic behavior and sociability that have carried us through the millennia and spread us into diverse environments across the planet. Clifford Geertz, of Princeton University, has described humans as “toolmaking, talking, symbolizing animals: Only they laugh; only they know when they will die; only they disdain to mate with family members; only they contrive those visions of other worlds called art. They have not just mentality but consciousness, not just needs but values, not just fears but conscience, not just a past but a history. Only they have culture” (1963, p. 2). Culture is a constellation of ideas and actions that are learned and transmitted from generation to generation. Human culture embodies the totality of behaviors and experiences that are summarized in our language and taught us by our parents and peers—a kind of group personality that provides a repertoire of actions in situations of choice. It is as impossible to have human identity without social contact as it is to have biological existence without parents. Tarzan was an ape until he met Jane. Culture enables us to eulogize our place in the universe, to create gods, to anticipate death, to travel to the stars—and to study archaeology.
extrasomatic Literally, “outside the
body”; nonbiological, nongenetic.
_ Images and Ideas Bones, Stones, and Human Behavior
69
pri31057_ch02.pp34-71.qxd
10/15/09
10:50 AM
p
Page 70
DISCUSSION QUESTIONS 1. What are the more important clues for “humanness” in the fossil record?
w w w. m h h e . c o m / p r i c e i p 6 e
For more review material and study questions, see the self-quizzes on your online learning center.
2. What is the significance of stone tools? Were these the first tools? 3. What were the four most important changes that took place as apes evolved into humans? 4. What can we say about the diet of the earliest humans? How did they obtain food? 5. Which of the sites and fossils discussed in this chapter seems to be the most important and why?
z
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e.
Aitken, M. J. 1990. Science-based dating in archaeology. New York: Longman. One of the best and more comprehensive overviews of physical dating methods used in archaeology, including radiopotassium and radiocarbon. Campbell, B. G., and J. D. Loy. 2000. Humankind emerging, 8th ed. New York: Longman. A well-illustrated textbook on human evolution and our early fossil ancestors. Cela-Conde, C. J., and F. J. Ayala. 2007. Human evolution: Trails from the past. New York: Oxford University Press. An up-to-date review of the evidence for human evolution. Fisher, H. E. 1983. The sex contract. New York: Quill. Another view of the origins of the species. Gamble, C. 2007. Origins and revolutions: Human identity in earliest prehistory. Cambridge University Press. Gilbert, W. H., and B. Asfaw. 2009. Homo erectus: Pleistocene evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press.
_ _
70
Chapter 2 The First Humans
Lewin, R. 1998. Principles of human evolution. Oxford: Blackwell. A popular and well-illustrated book on early humans. Stringer, C., and P. Andrews. 2005. The complete world of human evolution. London: Thames & Hudson. The latest in color on fossil humans. Tattersall, I., and J. H. Schwartz. 2000. Extinct humans. Boulder, CO: Westview Press. A book arguing that there have been many species of humans in the past and portraying their evolution. Taylor, R. E., and M. J. Aitken, eds. 1997. Chronometric dating in archaeology. New York: Plenum. A scientific overview of a variety of absolute dating methods. Zimmer, C. 2001. Evolution: The triumph of an idea. New York: HarperCollins. An exceptionally good book on the idea of evolution.
pri31057_ch02.pp34-71.qxd
10/1/09
7:48 AM
Page 71
_
pri31057_ch03.pp72-103.qxd
C
10/1/09
H
7:51 AM
A
Page 72
P
T
3 E
R
T
H
R
E
E
_ _ Figure 3.1 Reconstruction of a woolly mammoth at the Royal British Columbia Museum in Victoria.
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 73
Out of Africa Homo erectus
Introduction From Hominin to Human Our ancestors begin to discover Asia and Europe 2 million years ago
T
he magnificent creature shown in Figure 3.1 roamed the cold plains of Europe, Asia, and North America during the Pleistocene period. The woolly mammoth, a cold-adapted variety of elephant, typifies some of the adaptive challenges that emerged as the chill of the Pleistocene gripped the temperate zone, forcing plants, animals, and humans to change their location, appearance, and/or behaviors. Our first several million years were spent in Africa. The earliest dates for the presence of humans outside Africa are less than 2 m.y.a., the beginning of the Pleistocene. Hominins began to move into the more northerly continents of Asia and Europe and to encounter new environmental conditions. The Pleistocene, also known as the Ice Age, was a time of climatic extremes in many parts of the world. Repeated, dramatic changes in temperature, sea level, and environment are hallmarks of this time (see “Climate and Environment in the Pleistocene,” p. 81). The harsh environments of the Pleistocene Old World were the places we became more human, changing our habits, technology, and biology to adapt to these new conditions. Our hands and simple tools had been sufficient to obtain the foods available in the benign warmth of Africa. But expansion out of the tropics at the beginning of the Pleistocene required new skills and inventions for surviving where cold weather and the lack of food or shelter could be fatal. It became necessary for our ancestors to begin to change nature to fit their needs and enhance their survival. The first reliable evidence for the controlled use of fire, for systematic hunting, and for the use of wooden spears appeared during this time. Stone tools began to change as well. This is the period of the Lower Paleolithic, when core and flake tools replaced the pebble tools of the Basal Paleolithic. The handaxe was invented as an all-purpose tool (see “The Acheulean Handaxe,” p. 99) and is a hallmark of the Lower Paleolithic. The early migrants from Africa almost certainly belonged to Homo erectus, a new species of hominin, evolved from Homo habilis in Africa. Some researchers now separate the larger group of erectus into two distinct species, using the term Homo ergaster for early African varieties, and Homo erectus for later populations mainly in Asia and Europe. Fully modern humans first appeared in Africa some 200,000 years ago and likely evolved from Homo ergaster ancestors. For the sake of simplicity, we will use the general term erectus for this entire group of fossil humans.
73
< ( : < V : y < p : W
CHAPTER OUTLINE
Introduction From Hominin to Human 73 CONCEPT The Paleolithic Period 78 CONCEPT Climate and Environment in the Pleistocene 81 SITE Zhoukoudian 85 CONCEPT The First Europeans 89 SITE Atapuerca 91 CONCEPT Pleistocene Mammals 95 SITE Kalambo Falls and Olorgesailie 98 CONCEPT The Acheulean Handaxe 99
Images and Ideas The End of the Lower Paleolithic 101
w w w. m h h e . c o m / p r i c e i p 6 e
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
_
10/1/09
7:51 AM
Figure 3.2 Cranial capacity over time for Homo erectus. Two patterns emerge: (1) There is relatively little change between 1.5 and 0.5 m.y.a., and (2) there is a dramatic increase in cranial capacity after 0.5 m.y.a., perhaps associated with the rise of modern Homo sapiens.
Page 74
1400 1300
Cranial capacity (cc)
pri31057_ch03.pp72-103.qxd
1200 1100 1000 900 800 700
1.6 1.5 1.4 1.3 1.2 1.1 1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 Millions of years ago
TABLE 3.1 Major Characteristics of Homo erectus vs. Modern Humans
_ _
Figure 3.3 The Nariokotome skeleton, a 1.6-million-year-old fossil of a 12-year-old Homo erectus (also known as Homo ergaster) boy from Kenya. One of the most complete fossil skeletons from the early Pleistocene.
74
Trait
Homo erectus
Homo sapiens sapiens
Forehead Face Cranial capacity Lower jaw Teeth Brow ridges Limb bones
Absent In front of cranium 900 cc Larger and heavier; no chin Larger Heavy, across the eyes Larger and heavier
Vertical and rounded Under cranium 1350 cc Smaller and lighter; distinct chin Smaller Absent Smaller and lighter
The timing of the spread of populations out of Africa coincides closely with the appearance of H. erectus. The earliest H. erectus fossil comes from the western shore of Lake Turkana in northern Kenya, dating to approximately 1.8 m.y.a. H. erectus individuals were robust, with large bones and teeth; they also had larger bodies, more or less modern in size, and significantly larger brains— around 1000 cc—than their H. habilis ancestors (Table 3.1). These hominins were almost fully modern in movement and locomotion; they differed very little, if at all, from our own anatomy below the neck. Although their brain was twice as big as a chimpanzee’s, it was only about the size of a 1-year-old modern human child’s. H. erectus skulls are characterized by a low, sloping forehead; prominent brow ridges; and a protruding face. Cranial capacity changed very little in H. erectus during the period 1.8–0.5 m.y.a., after which a dramatic increase can be seen, perhaps in association with the rise of Homo sapiens (Figure 3.2). The trends in the development of Homo erectus show a number of similarities with modern Homo sapiens, including stature, increasing brain size, decreasing tooth and jaw size, and a vertical reduction in the size of the face. One of the more important erectus fossils is the so-called Nariokotome boy, discovered by Richard Leakey and Alan Walker in Kenya in 1984 (Figure 3.3). Among the most complete early human skeletons, the 10- or 12-year-old boy lived approximately 1.6 m.y.a. The skeleton resembled a robust modern one below the neck. Most of the differences between erectus and modern are seen in the skull. The Nariokotome skull exhibited a cranial capacity of 880 cc and the earliest evidence for a projecting, external nose.
Chapter 3 Out of Africa: Homo erectus
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 75
Figure 3.4 A hypothetical reconstruction of the face of Homo erectus.
Several sites in Asia document the arrival of the first migrants to that continent. The site of Dmanisi (dim-an-EASE-see) in the country of Georgia, containing an H. erectus lower jaw and other skeletal remains, has been dated by radiopotassium methods to 1.7 m.y.a. ’Ubeidiya (UB-a-de-ya), located in Israel a few kilometers south of the Sea of Galilee, dates to approximately 1.5 m.y.a. ’Ubeidiya contains concentrations of handaxes and other stone tools, along with animal bones, on the shore of a former lake. In this area, geological activity has tilted the layers so that the archaeological deposits are almost vertical. Even earlier dates are known from Central and East Asia. Locations of H. erectus finds on Java, one of the islands of Indonesia, also have recently been dated using radiopotassium methods to roughly 1.2 m.y.a. A recent find from Longgupo in southcentral China may go back even further. A fragment of a hominin jaw has been tentatively dated to 1.9 m.y.a. If these early dates prove to be accurate, Homo erectus groups may have spread very quickly across Asia following their initial appearance in Africa. Even at a slow rate of expansion of, say, 16 km (10 mi) per generation of 20 years, it would take only about 20,000 years to cover the distance between East Africa and Southeast Asia—around 16,000 km (10,000 mi). That brief period is indistinguishable by current chronology techniques such as radiopotassium dating. The date of the movement of human populations into Europe is another question (see “The First Europeans,” p. 89). The earliest dates for humans in Europe are just over 1 m.y.a., substantially later than the evidence from Asia. Questions about the origins of and differences in human skin color are common. A recent study by Alan Rogers of the University of Utah provides some insight on this issue. Based on mutation rates in modern genes for skin color, Rogers and his colleagues calculated that humans have been relatively free of body hair for 1.2 million years. One of the characteristics of being human is a relative absence of body hair compared to the other apes. The assumption is that our hominin ancestors, like chimpanzees today, originally had fair skin and dark hair (Figure 3.4). Rogers and colleagues’ thinking was that if humans have extensive body hair, there is no need for very dark skin color, which protects one from the harmful ultraviolet rays of the sun. Thus, the loss of body hair would have required that early humans in Africa develop dark skin color as protection against the sun. Homo erectus was eventually replaced by Homo sapiens after 100,000 years ago. Most of our knowledge about the period between 700,000 and 120,000 years
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the Nariokotome boy, see the Internet exercises on your online learning center.
Why did some of our early human ancestors leave the cradle of Africa?
Introduction From Hominin to Human
y _
75
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 76
ago, known as the Middle Pleistocene, comes from Europe and the Near East. Caves and rock shelters in those areas have preserved remains, attracting archaeologists interested in this period. The precise chronology for this important period in human evolution is unclear. Because the Middle Pleistocene is too early for radiopotassium methods of determining ages and too late for radiocarbon methods, accurate dates are rare. Many archaeological sites from this period have barely survived the elements and the passage of time. The sites described in this chapter document the discovery of these early humans and their activities (Figure 3.5). Zhoukoudian witnesses the presence of Homo erectus in East Asia. Atapuerca provides remarkable information on the first humans in Europe. Kalambo Falls and Olorgesailie demonstrate the continued development of the human species in Africa. These places take us through nearly 2 million years of human prehistory along our journey through time.
_ _
76
Chapter 3 Out of Africa: Homo erectus
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 77
Bilzingsleben Atapuerca
Terra Amata
Gran Dolina
Dmanisi Zhoukoudian
Isernia
Torralba
’Ubeidiya
Longgupo
Gesher Benot Ya’aqov
PACIFIC OCEAN
Lake Turkana ATLANTIC
Kalambo Falls
Olorgesailie
OCEAN
Kalambo Falls Equator
INDIAN OCEAN
Approximate distribution of Homo erectus
Sites discussed in the chapter
Lake Turkana Nariokotome Olorgesailie Kalambo Falls ’Ubeidiya Dmanisi
Gesher Benot Ya’aqov Zhoukoudian
Longgupo
Atapuerca
Atapuerca Torralba Terra Amata
2000
1500
1000
500
0
Thousands of years ago
_
Figure 3.5 Location of and timeline for Homo erectus in Africa, Europe, and Asia. Introduction From Hominin to Human
77
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 78
Concept
<
The Paleolithic Period Archaeological divisions of time elsewhere in Europe to designate the sequential ages of prehistory: Stone Age, Bronze Age, and Iron Age. These major divisions are still used in Europe and other areas of the world. The Stone Age was further divided in 1865 by the English naturalist John Lubbock, who coined the terms Paleolithic and Neolithic to distinguish the Old Stone Age and the New Stone Age, respectively. The Paleolithic is characterized by tools of flaked flint (Figure 3.7), and the Neolithic is represented by polished stone tools and pottery. Further divisions of the Paleolithic were made as the antiquity and the complexity of the period were realized. In 1872, the French prehistorian Gabriel de Mortillet proposed three major subdivisions of the Paleolithic: Lower, Middle, and Upper.
Just as geologists divide the history of the earth into periods and epochs, archaeologists break up the prehistory of human society into smaller, more manageable and understandable units. The scheme used by archaeologists to compartmentalize prehistory focuses on changes in artifacts and material culture. Differences in the types of material used to make tools and changes in the shapes of tools are often the main criteria for distinguishing time periods (Figure 3.6). The basic framework for dividing up the past was developed in 1836, when Christian Thomsen proposed an innovative three-age system for organizing the exhibits in the National Museum of Denmark, with separate display rooms for objects of stone, bronze, and iron. This system was quickly adopted
Paleolithic The first period of human prehistory, extending from the time of the first tools, more than 2.5 m.y.a., until the end of the Pleistocene, 10,000 years ago.
_ _
Figure 3.6 Tools of the Paleolithic. 1–3: handaxes from the Lower Paleolithic. 4–6: side scraper, denticulated blade, and Levallois point from the Middle Paleolithic. 7–18: Upper Paleolithic tools. 7: flint point. 8: end scraper on blade. 9: barbed bone harpoon. 10–11: flint points. 12: barbed antler harpoon. 13: backed flint knife. 14: flint point. 15: bone point. 16: blade borer. 17: bone needle. 18: bone point. For scale, the handaxe at the top of the photo is approximately 20 cm (8 in) in length.
Neolithic The period of time of early farmers with domesticated plants and animals, polished stone tools, permanent villages, and often pottery.
78
Chapter 3 Out of Africa: Homo erectus
2 1
3
4 5
6
9 8
7 10
12
11 18
13
15 14
17 16
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 79
An even earlier subdivision of the Paleolithic has been used for the earliest stone artifacts discovered in Africa. The Basal Paleolithic includes the pebble and flake tools of the Oldowan industry dating from around 2.6 m.y.a. until the appearance of handaxes. The Lower Paleolithic includes the Acheulean (ash-oo-LEE-an) assemblages generally associated with Homo erectus. Handaxes and flake tools characterize this time period (see “The Acheulean Handaxe,” p. 99). The Lower Paleolithic thus extends from approximately 1.9 m.y.a. to the beginning of the Middle Paleolithic, about 200,000 years ago. The Middle Paleolithic is associated with Neanderthals and other forms of early Homo sapiens and is characterized by a predominance of flake tools in artifact assemblages. In Europe and Southwest Asia, Middle Paleolithic assemblages are known as Mousterian. The Upper Paleolithic begins around 40,000 years ago with an emphasis on tools made of long, thin flakes of stone, known as blades, and on tools made from a number of other materials, including bone and antler (see “The Upper Paleolithic,” Chapter 4, p. 123). The finale of the Upper Paleolithic generally coincides with the end of the Pleistocene, about 10,000 years ago. Some of the major developments in the Paleolithic include the appearance of the first stone tools around 2.6 m.y.a., the controlled use of fire around 800,000 years ago, evidence for the hunting of large game by around 500,000 years ago, the first definite living structures perhaps 200,000 years ago, the intentional burial of the dead perhaps 100,000 years ago, the first art and decoration after 50,000 years ago, and the dispersal of human populations throughout the world by the end of the Pleistocene (Figure 3.8). Our human ancestors lived as huntergatherers throughout the Paleolithic, more than 99% of prehistory, successfully harvesting the wild foods of the land. Domestication—the planting of crops and the herding of animals—did
Figure 3.7 A major trend through the Paleolithic: increasing efficiency in the production of cutting edge. Pebble tools, handaxes, flakes, and blades were likely produced from the same original piece of flint. Blade production provides an enormous increase in the amount of cutting edge available from the same amount of material.
0.5 kg of flint
From 0.5 kg of flint: The pebble tool had 8 cm of cutting edge.
The handaxe had about 30 cm of cutting edge.
Mousterian flake tools provided about 90 cm of cutting edge. Upper Paleolithic blade production resulted in up to 9 m of cutting edge.
not begin until the very end of the Paleolithic, around 8000 B.C. Several important technological trends occurred during the Paleolithic, one of which was the increasing specialization of tools. The earliest stone artifacts were general-purpose tools— pebble tools, extremely simple in form. Over time, there was an increase in the kinds of tools and in the total number of kinds of tools in use. Efficiency in using stone also increased, as did the amount of cutting edge produced by flaking stone. For example, 0.5 kg (1 lb) of flint would produce about 8 cm (3 in) of cutting edge on an Oldowan pebble tool, about 30 cm (12 in) around the circumference on a handaxe from the Lower Paleolithic, and about 90 cm (30 in) of edge on the flake tools of the Middle Paleolithic. In the Upper Paleolithic, production of long, thin blades would result in almost 9 m (30 ft) of cutting edge. During the Paleolithic, there was also an increase in the variety of materials used to make tools. Bone, antler,
What relationship, if any, exists between brain size and human technology?
y
_ Concept The Paleolithic Period
79
pri31057_ch03.pp72-103.qxd
Years B.P.
Upper Pleistocene
Postglacial
10/1/09
7:51 AM
Climate Warmer Colder
10,000
100,000
Page 80
Period
Stone Industry
Archaeological Sites
Hominin Species
Major Events Farming
Neolithic Upper Paleolithic
Blade tools
Middle Paleolithic
Mousterian flake tools
Art
Lascaux, Pincevent Dolni Vestonice
Homo sapiens sapiens Homo sapiens neanderthalensis
Klasies River Kalambo Falls
Burial of dead Oldest dwellings
Middle Pleistocene
200,000 Terra Amata Olorgesailie Sima de los Huesos
Lower Pleistocene
500,000
Lower Paleolithic
Zhoukoudian Clactonian chopping tools
1,000,000
2,000,000
Acheulean handaxes
Basal Paleolithic
Oldowan pebble tools
Gesher Benot Ya’Aqov Gran Dolina Atapuerca Dmanisi ’Ubeidiya
Use of fire Spread out of Africa Homo erectus Handaxes
Koobi Fora Olduvai Swartkrans
Homo habilis
Large brains
Pliocene
First stone tools Hadar
3,000,000
Laetoli Australopithecines 4,000,000 No Archaeology
Middle Awash
5,000,000
Ardipithecus
Oldest hominid fossils
6,000,000
Figure 3.8 The Pleistocene and the Paleolithic. This chart shows the major divisions of the Pleistocene and years before present, along with the major divisions of the Paleolithic, important stone industries, archaeological sites, hominin forms, and significant events.
I tend to live in the past because most of my life is there. —Herb Caen
_ _
80
ivory, and wood were commonly used by the end of the period, although that evidence may be a result of better preservation at the more recent sites. The Paleolithic witnessed the achievement of humanness, a heritage that has been passed on to the inhabitants of the most recent 10,000 years of our
Chapter 3 Out of Africa: Homo erectus
species’ past. The major developments in that recent past would not have been possible without the population expansion, the innovative technology, and the development of language, social relationships, and ritual that characterized the journey of our ancestors through the Paleolithic.
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 81
Concept
<
Climate and Environment in the Pleistocene Challenging conditions for human survival with the temperature of the ocean water or the atmosphere and can thus indicate temperature change over time. The amount of the 18O isotope relative to the 16O isotope decreases with increasing temperature. For the Pleistocene, alternations between warmer and cooler periods defined by oxygen isotopes have been numbered sequentially and are defined as isotope stages (Figure 3.10). These stages are sometimes used for the chronology of the Paleolithic. The figure shows these years before present, changes in isotopes and temperature, stage numbers, and divisions of the Pleistocene for the past 800,000 years. It is not difficult to imagine the consequences of a sharp drop in temperature during the colder episodes of the Pleistocene. In parts of northern Canada today, temperatures are cool even in summer, with lows at night around 4.5°C (40°F). The last snowfall is often in June, and the first flakes of autumn come in September. A drop in average summer temperature in this area of about 5°C (8°F) would mean that not all the winter snow would
Cretaceous
Paleocene
Eocene
Figure 3.9 Oxygen isotope ratios, a proxy for atmospheric temperature, from foraminifera in marine sediments during the Cenozoic, the past 70 million years of the earth’s history. Notice the gradual decline in atmospheric temperature since the Eocene.
Oligocene
Plio/ Pleistocene
Miocene
–1.0 Oxygen isotope ratio
The Pleistocene, a geological epoch and the time of Homo erectus, began approximately 2 m.y.a. in a period characterized by active volcanoes, cooling temperatures, and the appearance of several modern species of animals. The earth’s temperature had gradually been cooling for millions of years (Figure 3.9). Although the Pleistocene is also known as the Ice Age, there were, in fact, many ice ages, or glacials, as they are known. There were at least nine, and probably more, alternations between colder and warmer conditions, known as glacials and interglacials, during the Pleistocene. These glacial episodes were first recognized in Europe. During the 1800s, two Swiss geologists, A. Penck and E. Brückner, identified four periods of glaciation in Europe from a series of Alpine deposits and river terraces. They named these glacial stages after local rivers—the Günz, the Mindel, the Riss, and the Würm—from oldest to most recent. The intervening warm periods were designated Günz/Mindel, Mindel/Riss, and Riss/Würm. However, subsequent research has demonstrated that in the past million years or so, there have been a larger number of pronounced swings from warmer to colder. It may be surprising to learn that information about the dramatic changes in temperature and glaciation in the Pleistocene has come from underwater in the Caribbean Sea and from the ice cap in Greenland, where scientists have made deep corings of marine sediments and glacial ice, respectively. The small animal shells (foraminifera) that make up much of the sediment on the ocean floor contain oxygen isotopes in the carbonate of the shell. In Greenland, oxygen isotopes from the Pleistocene are preserved in the kilometer-thick layer of glacial ice that covers the huge island. Oxygen isotope ratios vary
0.0 1.0 2.0 3.0 70
60
50
40
30
20
10
0
_
Millions of years ago
Concept Climate and Environment in the Pleistocene
81
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 82
Oxygen isotope ratio Years
Colder
B.P.
Isotope stage 1 2 3 4 5
13,000 32,000 64,000 78,000
Upper Pleistocene
128,000 6 195,000
7
251,000
200,000
8
297,000
9
347,000 367,000
10 11
440,000 472,000 502,000
12
Middle Pleistocene
13 14
542,000
15
592,000 627,000 647,000 688,000 706,000 729,000
16 17 18 19 20 21
780,000
782,000
Lower Pleistocene Odd-numbered stages (in blue) = warmer periods, less glacial ice cover Even-numbered stages = colder periods, more glacial ice cover
Figure 3.10 Pleistocene glacial chronology based on oxygen isotope ratios in foraminifera in marine sediments. Notice the many oscillations between cold and warm phases over the past 800,000 years or so. (Data from Shackleton and Opdyke, 1973.)
glacial A cold episode of the Pleis-
tocene; also called an ice age. interglacial A warm period of the
Pleistocene. glaciation The expansion of continental glacial ice during a period of cold climate. oxygen isotope ratio The ratio of
_ _
different isotopes of oxygen in ocean water, varying with the temperature of the water.
82
melt each year; snow would fall later in summer and earlier in autumn. Even if these snowbanks increased by only a few centimeters each summer, they would grow substantially over the course of time. For example, just 10 cm (4 in) of snow every year for 10,000 years would build an enormous snowdrift over 1 km (0.6 mi) high. The 5 trillion snowflakes it takes to cover a football field with a foot of wet snow weigh almost 550 tons. Ice forms under piles of accumulating snow. Its weight causes the mass of ice to begin to spread at the edges. In such a manner, huge sheets of ice expanded horizontally during the Pleistocene, as more and more snow and ice accumu-
Chapter 3 Out of Africa: Homo erectus
lated. It has been estimated that such ice sheets expanded across the landscape at a rate of as much as 100–150 m (350–500 ft) per year. In a thousand years, such a growing ice mass would cover a distance of 100 km (62 mi). In 20,000 years of accumulation, the ice could have spread 2000 km (1250 mi) from its original center. Today, perhaps 10% of the land surface of the earth is covered by glaciers—in Antarctica; Greenland; northern Canada, Europe, and Asia; and high mountain regions. During the past million years, sheets of continental glaciers more than 5 km (3 mi) thick grew in the Northern Hemisphere. During the colder episodes of the Pleistocene, continental ice sheets expanded from these same areas to cover perhaps 30% of the land surface (Figure 3.11). In North America, these ice masses filled most of Canada and extended into the United States as far south as what is now St. Louis, Missouri. In northern Europe, a similar sheet moved from the Baltic Sea basin to cover Scandinavia, northern Great Britain, parts of the Netherlands, northern Germany, Poland, and the northwestern part of the former Soviet Union. These sheets of ice acted like enormous bulldozers, grinding down the landscape as they advanced and depositing huge blankets of homogenized earth and rock, known as moraines, as they retreated. The weight of the ice also forced down the land surface, often to great depths. In Greenland today, for example, where the ice sheet is 2 km (1.3 mi) thick, the land surface resembles a very deep bowl, higher at the edges, where the ice is thinner. The land surface in the interior of Greenland is far below present sea level and one of the lowest spots on earth. The masses of ice reduced the earth’s liquid water reservoir, particularly the oceans. During the time of maximum cold, when water was frozen in huge continental sheets, the global sea level was lowered as much as 150 m (500 ft), completely changing the outlines of the continents and often creating connections between former islands and separate landmasses.
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 83
As the massive ice sheets melted at the end of the Pleistocene, two processes operated to change the shape of the continents. At the same time the melting ice refilled the seas and coastlines moved higher, the enormous weight of the ice sheets disappeared, removing a heavy burden from the land. The surface of the earth had been depressed by the weight of the ice (just as the center of Greenland is today). This surface began to come back, or rebound, as the ice disappeared. For example, the coast of Finland today is expanding at the astounding rate of several meters per century as the area recovers from the weight of the glacial ice. The causes of Pleistocene climatic change and continental glaciation have been debated for many years. Volcanic plumes filling the air with ash and smoke and shading the earth or changes in the level of solar radiation were once thought to be responsible for the onset of the Pleistocene. Recent attention has been given to the role of
mountain uplift in cooling the earth’s climate. Walter Ruddiman and John Kutzbach have suggested that the uplift of the North American Rocky Mountains and the Himalayas of South Asia resulted in a global disruption of weather patterns and the onset of a cooler, drier climate. Fluctuations in air and water temperature were not uncommon during the Pleistocene and are cyclical in nature. In the 1920s, Serbian mathematician Milutin Milankovitch argued that the variations in the earth’s orbit changed climate in a cyclical fashion. Slight variation in the precession of the earth’s axis and shifts in orbit eccentricity change the distribution and intensity of sunlight reaching the planet, like the changes in our seasons. Milankovitch predicted cycles of 100,000, 40,000, and 20,000 years for these climatic changes, based on his calculations of the earth’s orbital variation. Such a cyclical pattern seems to fit the information on climatic change found in the oxygen isotope cores.
Figure 3.11 The maximum distribution of continental glaciation during the colder periods of the Pleistocene. Ice sheets are more common in the Northern Hemisphere partly because there is much more land at the higher northern latitudes.
Concept Climate and Environment in the Pleistocene
_
83
pri31057_ch03.pp72-103.qxd
10/1/09
Milankovitch forcing A term de-
scribing the phenomenon considered to be the prime reason for glacial fluctuations and climatic change.
7:51 AM
Page 84
Thus, it appears that minor changes in factors such as the distance from the earth to the sun and the tilt of the earth’s axis play a major role in the amount of sunlight reaching the earth, the atmospheric temperature, and ultimately the expansion and retreat of continental glaciation. This factor, now known as Milankovitch forcing, is considered to be the prime reason for the cycle of fluctuating temperatures that marks the Pleistocene. It is not at all clear that the glacial cycles of the Pleistocene have ended; we may now simply be witnessing one of the warmer
_ _
84
Chapter 3 Out of Africa: Homo erectus
intervals between fluctuations of glaciation and climatic change that occurred during the Pleistocene and repeatedly over the past several hundred million years of the earth’s history. The end of the Pleistocene is dated to 10,000 years ago (8000 B.C.), when the last ice sheets retreated. The interval that we presently occupy is known as the Holocene (a.k.a. the Postglacial, or the Present Interglacial). The similarity of our present interglacial to an earlier one about 400,000 years ago suggests that the next glacial will begin in about 20,000 years.
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 85
Site
:
Zhoukoudian Bones of the dragon
For millennia, many Chinese believed that fossil bones had medicinal and curative powers. Called dragon’s teeth, such fossils were ground into powder and sold at apothecaries throughout the country. For more than a century, paleontologists and other natural scientists have visited such shops to look for the bones of new species and to learn about potential new fossil sites. In 1899, a European doctor in Beijing found an unusual fossil tooth at one such apothecary and identified it as an upper third molar of either human or ape origin. The tooth came from a place called Dragon Bone Hill, a large limestone ridge near the town of Zhoukoudian (joe-ko-tea-EN), 50 km (30 mi) southwest of Beijing (Figure 3.12). Before the doctor returned to Europe, he passed the tooth and the information on to a Swedish geologist,
John Gunnar Andersson. Andersson and his friend Davidson Black, a Canadian anatomist, were convinced they could discover an early human fossil where the tooth had been found. Black persuaded the Rockefeller Foundation to sponsor excavations at the site. On the basis of two hominin teeth that were found, Black announced the discovery of Sinanthropus pekinensis (Chinese man of Peking) in 1927. Later that same year, the first skull was found, confirming Black’s bold proclamation. For 10 years, a large workforce essentially mined the deposits in the complex of caves at Zhoukoudian, removing over half a million tons of material in the quest for fossils (Figure 3.13). Almost 2000 days, more than 6 months each year, were spent blasting out the limestone and removing rock and sedimentary deposits over a
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on Zhoukoudian, see the Internet exercises on your online learning center.
Figure 3.12 The location of the excavations at Dragon Bone Hill, Zhoukoudian.
_ Site Zhoukoudian
85
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Figure 3.13 Excavations in the lower levels at Zhoukoudian. An enormous amount of material was removed from the deposits at the site, the volume equivalent in size to a small building.
_ _
Figure 3.14 Entrance to the site of Zhoukoudian today.
86
Page 86
vertical distance of 55 m (180 ft), as high as a 17-story building. The large limestone chamber at the center of the Zhoukoudian caves is enormous, 140 m (450 ft) north to south, by 40 m (125 ft) east to west, by approximately 40 m (125 ft) high—the size of a supertanker. The deposits in this chamber were almost completely removed in the course of the excavation (Figure 3.14). The crude excavation methods and untrained labor meant that stone tools and other materials were often missed and that important information about the context of the deposits was not recorded. In his report, Black described dense layers of ash, baked sediments, and charred bone resulting from fires in the cave (Figure 3.15). These materials were thought to be evidence of the places where people lived, made tools, built fires, ate, died, and left their
Chapter 3 Out of Africa: Homo erectus
bones. Over 20,000 stone tools, including flakes, scrapers, and choppers (but no handaxes), were found, made from quartz, sandstone, rock crystal, and flint (Figure 3.16). These materials do not occur naturally in limestone areas and must have been brought into the caves. The artifacts are generally very crude and irregular but do improve in quality toward the top of the deposits. The abundant bones in the deposits come from both large and small species. Most of the large animal bones come from an extinct species of deer with enormous horns and from wild horses and giant boars, elephants, water buffaloes, hyenas, such carnivores as bears and saber-toothed tigers, and others—a total of 96 mammalian species. The presence of these animals indicates that the climate was somewhat warmer than it is today. Moreover,
pri31057_ch03.pp72-103.qxd
10/1/09
Geological Period
7:51 AM
Page 87
Stratigraphy
Layer
5m
Breccia travertine
10 m
Breccia with ashes
15 m
Ashes
Figure 3.15 The stratigraphy of 40 m (130 ft) of deposits at Zhoukoudian, showing the location of the Homo erectus skeletal remains.
Hard travertine 20 m Hard breccia
Middle
Middle Pleistocene
Upper
Depth
25 m
Fine sand
30 m
Breccia with ashes
Upper red clay Lower ash
35 m Lower
Breccia Coarse sand Reddish clay sand
40 m Lower Pleistocene
Basal gravel
Homo erectus fossils
the habitat requirements for these species suggest that the area around Zhoukoudian was a mosaic of forested hills, open plains, lakes, and rivers. The forest was likely dominated by pine, cedar, elm, hackberry, and the Chinese redbud tree. The charred seeds of hackberry fruits found in the deposits at Zhoukoudian led to the suggestion that plant foods such as these may have formed part of the diet of the human inhabitants. By far the most important finds at Zhoukoudian were the remains of the early hominins, today designated Homo erectus. A total of 6 skullcaps (the face and the lower portion of the crania are
missing), 12 skull fragments, 15 lower jaws, 157 teeth, 7 thighbones, 1 fragment of shinbone, 3 bones from the upper arm, 1 collarbone, and 1 wrist bone were recovered. As in the South African caves, hominin skulls are more common than other bones of the skeleton, partly because they are more resistant to destruction and partly because they are more readily recognized. These fragments come from adult males and females and from children. An increase in brain size over time can be seen in the materials from Zhoukoudian. Skulls from the deeper part of the deposits have a cranial
_ Site Zhoukoudian
87
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 88
Figure 3.16 Stone artifacts found at Zhoukoudian. These materials were largely flakes and small chopping tools. No handaxes were recorded at this site.
Burin
Points
Scrapers
capacity of about 900 cc, whereas those from the upper levels are closer to 1100 cc on average, within the range of variation of modern humans. The investigations conducted by Black, and since 1949 by Chinese archaeologists, have revealed 12 stratified layers in the deep deposits at Zhoukoudian (see Figure 3.15). The layers themselves date to at least 700,000 years ago at the bottom and to roughly 200,000 years ago at the top of the sequence. Unfortunately, these important fossils have disappeared. The excavations at Zhoukoudian were closed in 1937 with the outbreak of war between Japan and China. A decision was made to move the fossils to the United States
_ _
88
Chapter 3 Out of Africa: Homo erectus
Chopper
for safekeeping. They were packed carefully in crates and placed in the hands of a detachment of U.S. Marines. Somewhere between the consulate in Beijing and the port of departure from China, the fossils were lost or stolen. Today they may be still in China, in Japan, in the United States, or at the bottom of the Pacific—no one knows. Fortunately, plaster casts were made before the fossils were lost, so that there is at least some information on their size, shape, and important features. In addition, the more recent Chinese excavations since 1949 have uncovered a few more examples. Nevertheless, these priceless relics of some of our earliest ancestors would be a marvelous rediscovery.
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 89
Concept
<
The First Europeans An early date of arrival
As we have seen, early hominins began to migrate out of Africa into Asia at the beginning of the Pleistocene, by at least 1.9 m.y.a. The earliest evidence for humans in Europe is now 1.2 million years old. If Homo erectus spread into Asia, why did they not move into Europe at the same time? The answer to this question may simply be that we don’t yet have enough information, or it may relate to the geography of the European continent. New York City lies at the same latitude as Madrid and Rome. In fact, much of Europe is farther north than the U.S. border with Canada (Figure 3.17). Very cold climatic conditions would have prevailed in Europe during much of the last 1 million years of the Pleistocene, making large portions of the continent generally inhospitable to human occupation. Huge glaciers extended across the Pyrenees, the Alps, Scandinavia, and Britain during the Pleistocene. Permafrost tundra covered much of the central part of the continent during colder periods. Thus, the arctic and subarctic climate of Europe may have delayed human settlement until behavior and technology overcame this obstacle. There are several very old archaeological sites in Europe. The earliest is in northern Spain, a remarkable 1.2million-year-old site at Atapuerca, excavated since 1994 and described in the next section. Another early site has been found in Italy, at a place known as Isernia. A few simple stone tools and animal bones were found together in deeply buried lakeshore deposits underlying a volcanic ash that has been dated to 730,000 years ago using radiopotassium methods. Large animal species associated with the deposit are bison, rhinoceros, elephant, bear, hippo, pig, goat, and deer.
Other pieces of early humans have been found around the continent: a lower jaw from Mauer near Heidelberg and a crushed skull from Steinheim in Germany, a skull from Petralona in Greece, the facial front of a skull from Arago in France, and a skull from Swanscombe and a leg bone from Boxgrove in England (Figure 3.18). These European fossils date to the period after 600,000 years ago and appear slightly more modern than their Homo erectus neighbors in Asia. These remains have sometimes been referred to as Archaic Homo sapiens, but they are now designated as Homo heidelbergensis.
Why does large-game hunting play an important role in human survival?
y
Figure 3.17 The map of Europe shows the locations and ages (in thousands of years) of some of the earliest archaeological sites mentioned in the text. Europe is superimposed over the United States to show the relationship of the two continents in size and latitude.
Clacton (245)
Bilzingsleben (350) Swanscombe (325)
50° North
Steinheim (250) Terra Amata (350) Atapuerca (1200) Arago (250)
Verteszöllös (350)
Isernia (730)
Torralba and Ambrona (350)
_ Concept The First Europeans
89
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 90
Figure 3.18 Early hominin skulls in Europe: Steinheim (left) and Arago (right). Dating to 300,000–200,000 years B.P., these specimens likely represent an intermediate form between Homo erectus and fully modern Homo sapiens. Thus, they are now considered as Homo heidelbergensis and the ancestors of Homo neanderthalensis.
Figure 3.19 Excavations at the site of Ambrona in central Spain, dating to 350,000 years B.P. The site documents the association of stone tools and the bones of extinct animals such as the woodland elephant. Notice the linear arrangement of bones and tusks.
_ _
90
A few other sites in Europe may be more than half a million years old, but reliable dates are rare. By 400,000 years ago, however, most of Europe had been occupied. Lower Paleolithic sites are found throughout the central and southern parts of the continent at places such as Bilzingsleben in Germany, Verteszöllös in Hungary, and Torralba and Ambrona in Spain.
Chapter 3 Out of Africa: Homo erectus
Bilzingsleben lies in an area of natural mineral springs; human and animal bones and small stone tools were preserved in the hardened deposits of travertine around the springs. Isotopic dates for the site indicate an age between 400,000 and 300,000 years old, during an interglacial period. Human skull fragments were found in the deposits, along with numerous animal bones. A site very similar to Bilzingsleben exists in Hungary at Verteszöllös, roughly 50 km (30 mi) west of Budapest. Here again, the stone artifacts are curious because of their small size, probably the result of an absence of larger pieces of raw material in the area. Short cutting edges were produced on thousands of tiny pebbles, often only 2.5 cm (1 in) in diameter. Charcoal and burned animal bones document the use of fire at this site. As at Bilzingsleben, fragments of skull from early European hominins were found in the layers at Verteszöllös. Some 150 km (100 mi) northeast of Madrid, a steep-sided valley cuts through a high plateau, creating one of the few routes between north and south in the region. During the Pleistocene, this valley almost certainly would have been an important path of migration for large animals moving between the north in summer and the south in winter. Two almost-identical sites, Torralba and Ambrona, are located in this valley, containing deposits of stone artifacts and the bones of extinct elephants, horses, deer, and other animals (Figure 3.19). The bones and stones accumulated here during a colder episode of the Pleistocene, probably during a glacial period, perhaps 350,000 years ago. The artifacts at the sites include objects made of stone, bone, ivory tusk, and wood. The stone artifacts belong to the early Acheulean; handaxes were made and used here, along with cleavers and a variety of other flake and core tools. Some of the stone used for making these artifacts came from tens of kilometers away.
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 91
Site
:
Atapuerca Remains of the first Europeans, 1.2 million years B.P. vation and the need to maintain a reasonably level grade, several deep cuts, or trenches, were hacked through the limestone for the rail line. This railroad trench in fact exposed a number of caverns and chambers that had been hidden for millennia (Figure 3.20). It was not until the 1970s, however, that prehistoric artifacts and bones were first found in these deposits and not until the 1990s that major excavations began to expose the richness of their contents (Figure 3.21). There are many interesting layers in the Atapuerca hills, but the two most important for the Paleolithic are known as Sima de los Huesos (SEE-ma day lohs
The limestone hills of Atapuerca (ahtah-PWER-ka), near the city of Burgos in northern Spain, hold a treasure trove for Paleolithic archaeologists. Human remains and artifacts from the Lower, Middle, and Upper Paleolithic are hidden in the many caves of this porous rock mass. Limestone is a soft rock, gradually dissolved by rain and groundwater, which cut channels, caves, and underground streams through the hills. The entrances to many of these caverns collapsed over time, closing and hiding the chambers and entombing the materials within. Then, in the 1890s, a railroad was built through the hills of Atapuerca. Because of the ele-
FRANCE Gran Dolina PORTUGAL
Gran Dolina
Railroad cut
Atapuerca
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on Atapuerca, see the Internet exercises on your online learning center.
Figure 3.20 A plan of the railroad cut and some of the large caves and chambers exposed in the Atapuerca hills. The locations of Gran Dolina and Sima de los Huesos are shown on this plan.
Madrid SPA I N
N
Sima de los Huesos 0
100
200 m
_ Site Atapuerca
91
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 92
Figure 3.21 Excavations at Gran Dolina in the Atapuerca railroad cut. The scaffolding is necessary to reach the deposits high on the side of the trench wall.
y
Do artistic reconstructions of the faces of past ancestors help or hinder our knowledge of their appearance?
WHEY-soss) and Gran Dolina (graan doe-LEEN-ah). The remains from Gran Dolina are immediately relevant to our discussion of Homo erectus, but the finds from Sima de los Huesos are equally fascinating and will be described briefly before we turn to Gran Dolina.
Figure 3.22 An artist’s reconstruction of the 32 individuals placed in the “pit of bones” at Atapuerca, dating to more than 350,000 years ago.
_ _
92
Chapter 3 Out of Africa: Homo erectus
Sima de los Huesos or “pit of bones” is an extraordinary crevice deep in one of the largest caves in the Atapuerca region. The “pit” lies 55 m (175 ft) below ground and 500 m (1500 ft, or five football fields) from the nearest entrance to the cave. The pit is
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 93
Figure 3.23 The face of Homo erectus, one of the best-preserved skulls from Gran Dolina, dating to approximately 900,000 years ago—one of the earliest inhabitants of Europe.
“sock shaped,” a small depression at the end of a sloping passage entered from a 13-m (40-ft) vertical shaft. Merely reaching this crevice is a difficult journey for the excavators. The passage is narrow and slippery, and part of the trip requires traversing a deep shaft. Only one or two individuals can work in the space that exists in this nook of the cave. The excavations, however difficult, have produced spectacular results and the largest set of human remains ever recovered from the Paleolithic. Parts of the skeletons of at least 32 individuals have been found (Figure 3.22). Nine are male, and 9 female; the sex of 14 could not be determined. The largest number, 11 individuals, were between ages 13 and 17. Four children were between ages 3 and 13. Only 4 people were over age 30 when they died, and no one had reached age 40. These individuals were of normal height; males averaged 1.75 m (5 ft, 9 in), females 1.7 m (5 ft, 7 in). The tallest person was 1.8 m (5 ft, 11 in) in height and weighed 90
kg (200 lb). The cause of death is unknown at present. The remains are an extraordinary sample of a human group living during the Lower Paleolithic. What is most remarkable is that all these individuals were intentionally put into the pit. Their bodies were carried into the cave, through the difficult passage, and placed or dropped in the crevice. Perhaps this behavior foreshadows the practice of intentional burial of the dead that is first seen among Neanderthals in the Middle Paleolithic, some 300,000 years later. Uranium-series dating of redeposited limestone at Sima de los Huesos produced an estimated age between 350,000 and 500,000 years. The human remains in Sima de los Huesos have been designated as Homo heidelbergensis and represent the ancestors of the Neanderthals we will discuss in the next chapter. It is in regard to Homo erectus, however, that the finds from Gran Dolina are the most important and represent the oldest yet-discovered _ Site Atapuerca
93
pri31057_ch03.pp72-103.qxd
y
10/1/09
Does cannibalism indicate that there were food shortages in the Lower Paleolithic?
7:51 AM
Page 94
humans on the continent of Europe. It is intriguing that the earliest finds should be in the westernmost parts of Europe. The deposits at Gran Dolina originally accumulated at the entrance of a cave or rockshelter that has largely disappeared. The original entrance area was exposed when the deep railroad cut was made through the area. High scaffolding is necessary to reach the deposits where the excavations take place. There are many layers at Gran Dolina, but an important one is Layer 6, which has been dated by magnetic reversal to approximately 900,000 years ago. This layer is one of the richest in the remains of the bones of more than 25 animal species, such as horse, deer, bison, rhinoceros, wild cat, hyena, and wolf. Among the bones have been found several hundred stone tools, for the most part simple pebble tools and flakes along with a few handaxes. Among the stone and the animal bones were also found some 90 human fossil fragments, representing at least four individuals. These bones belong to a
_ _
94
Chapter 3 Out of Africa: Homo erectus
form of Homo erectus (called Homo antecessor by the Spanish scientists) (Figure 3.23). These bones exhibit a pattern of breakage and of cutting and chopping marks from stone tools that clearly indicates that the humans who butchered the animals represented by the bones in the cave also butchered and must have consumed the humans. Gran Dolina documents the fact that the early Europeans were cannibals. And the discoveries keep coming. In 2008, the archaeologists and paleoanthropologists working at Atapuerca announced a new and even older find. A portion of a human mandible, several stone tools, and broken animal bones were found in a deep layer in an area of the caves known as Sima del Elefante. Several methods were used to obtain an estimate of the date for this layer and the archaeological finds of 1.2 m.y.a., the oldest human remains yet found in Europe. Stay tuned; there will no doubt be more extraordinary discoveries from this remarkable site at Atapuerca.
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 95
Concept
<
Pleistocene Mammals Prehistoric species now extinct
Although the Pleistocene witnessed the appearance of a number of modern mammals (e.g., elephants, horses, and cattle), many other animals that existed then are now gone. In North America, a group of animals largely different from those we see today wandered across the cold and largely treeless regions—woolly mammoths, mastadons, ground sloths, caribou, giant beavers, saber-toothed tigers, bison, and horses (Figure 3.24). Most are now extinct. In Europe, woolly mammoths and woolly rhinoceroses, cave bears and cave lions and cave hyenas, giant deer and beavers, bison and aurochs (wild cattle) roamed over much of Europe during the colder periods of the Pleistocene. These species were generally much larger than their modern equivalents and are now extinct. The bones of these animals from
archaeological sites give some indication of the species that were present in different parts of Europe. The cave lion and the cave bear are generally well known because of the large number of their bones that have been found in the caverns and grottoes of Europe. The cave bear was enormous, at least the size of a modern Alaskan grizzly, with a very large head and huge canine teeth. The cave bear ranged over much of Europe, and many of this creature’s bones have been found in caves at high elevations in the Alps. Tight crevices in the caves had been polished by the passage of bears over tens of thousands of years. Thousands of skeletons of cave bears that died during hibernation have been preserved. The Drachenhöhle in Austria is estimated to have contained the remains of over 30,000 bears.
Figure 3.24 Notable among the mammals found in Pleistocene North America were beavers the size of a small bear, woolly mammoths, mastodons, sloths, saber-toothed tigers, bison, and horses. Many of these same species lived across the colder regions of Asia and Europe. Most are now extinct.
_ Concept Pleistocene Mammals
95
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 96
Figure 3.25 Excavation of a large bull aurochs skeleton at Prejlerup in eastern Denmark. This animal escaped from its hunters but died from its wounds.
The woolly mammoth is also well known, both from the skeletal material that has been found and from a few examples of almost complete animals— soft parts, skin, and hair—found frozen in the permafrost of Siberia and Alaska. In Alaska, mammoth hairs still frozen in the ground sometimes clog the equipment of gold miners. This animal had a huge domed head, enormous curved ivory tusks, and humped shoulders. Standing approximately 3–4 m (10–13 ft) at the shoulder, it was covered with a short, woolly undercoat and long, hairy overcoat. The mammoth inhabited arctic steppe environments and consumed a diet of tundra grasses. This species ranged from western Europe across northern Asia and into North America during the late Pleistocene. The giant Irish deer, from the bogs of Ireland and northern Europe, was known for its enormous antlers, as much as 4 m (13 ft) across. This animal was about the size of a North American moose, but it more closely resembled an elk. Wild cattle, known as aurochs, were common game animals during the warmer periods of the Pleistocene and survived until the middle of the nineteenth century A.D. (Figure 3.25). These animals were probably adapted
_ _
96
Chapter 3 Out of Africa: Homo erectus
to woodlands and were particularly common in the early Postglacial. A number of these species disappeared around the end of the Pleistocene (Figure 3.26); mammoths, cave bears, and many large species were extinct by 11,000 B.C. (see “Pleistocene Extinction,” Chapter 4, p. 151). The giant Irish deer may have survived around the Black Sea until 3000–2000 years ago, long after it had disappeared from the European continent. The last aurochs died in a game forest in Poland in A.D. 1627. The largeheaded Prezwalski horse, depicted in Paleolithic cave paintings, still exists, in several European zoos. Other important Pleistocene species, such as reindeer, red deer, elk, musk ox, and brown bear, survive to the present and continue to exist at the edges of urban civilization.
pri31057_ch03.pp72-103.qxd
10/1/09
Lagomorph
Page 97
Lion
Mosbach wolf
Macaque Ambrona, Upper (NISP = 1046) Ambrona, Lower (NISP = 2458) Torralba (NISP = 956)
7:51 AM
Spotted hyena
Straight-tusked woodland elephant
Lynx
Steppe rhino
Wild horse
Fallow deer
Red deer
Aurochs
Giant deer
Figure 3.26 Species of animals represented by the bones at the Spanish Lower Paleolithic sites of Torralba and Ambrona. The width of the bar for each site and level indicates the abundance of the species. Elephants and horses are by far the most common at these sites. “NISP” stands for the number of identifiable bones per species.
_ Concept Pleistocene Mammals
97
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 98
Site
:
Kalambo Falls and Olorgesailie East Africa 200,000 years ago
Figure 3.27 Wooden artifacts, including a club (left) and a sharpened object (right) from Kalambo Falls.
_ _
The Kalambo River flows peacefully through the plateau country of East Africa until its waters suddenly leap over the edge of the Rift Valley at the border of Tanzania and Zambia. There, the river falls over 250 m (800 ft) into a spectacular gorge close to the present shore of Lake Tanganyika. Deposits in the valley above the falls have provided one of the longest archaeological sequences anywhere in East Africa—or the world, for that matter. Over 200,000 years of successive human occupations are preserved in the water-lain deposits from the floods and backwaters of the river. The site of Kalambo Falls was first discovered in 1953 by Desmond Clark, of the University of California at Berkeley, while he was examining a steeply eroded bank of the river. Preserved tree trunks and branches were present in the layers at the bottom of the bank, along with handaxes and cleavers of the Acheulean tradition in almost fresh condition (Figure 3.27). This lowest level, designated Bed I, is over 3 m (10 ft) thick and composed of alternating layers of white sand and dark clay. Kalambo Falls is a very unusual site for two reasons. First, evidence for the
Figure 3.28 An Acheulean handaxe from Kalambo Falls shown in front view and cross section.
98
Chapter 3 Out of Africa: Homo erectus
use and consumption of plants is preserved at this early site; and second, evidence for meat-eating in the form of animal bones is not represented. The acidic sediments at the Kalambo site removed all traces of bone, but the waterlogged condition of the lower layers allowed the preservation of a variety of plant materials. Bed I contains several living floors from the late Acheulean period, along with a remarkable set of plant materials, including leaves, nuts, seeds, and fruits, and some of the oldest wooden objects in the world. A club, a smoothed and pointed piece, and other wooden objects shaped by human hands were recovered from the lowest Acheulean layers (Figure 3.28). Evidence for fire is also preserved in charred logs found on the living floors. In contrast to Kalambo Falls, the site of Olorgesailie (o-lorg-a-SIGH-lee) in Kenya documents meat-eating in the Acheulean of Africa. The site was located along a sandy streambed at the edge of a former lake and contained over 400 handaxes and numerous animal bones. Much of the stone raw material for the handaxes came from nearby Mount Olorgesailie. Over a ton of lithic material was transported to the site, testimony to the importance of stone tools for these Acheulean groups. Bone preservation was exceptionally good at Olorgesailie. Most of the animal bones come from at least 65 individuals of an extinct species of very large baboon, known as Theropithecus. All the skeletal elements of the baboons were represented, both large and small bones. There is strong evidence for one or more episodes of humans hunting baboons along the tree-lined stream banks at Olorgesailie. The two sites of Kalambo Falls and Olorgesailie together document the importance of both plants and animals in the diet of Homo erectus.
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 99
Concept
<
The Acheulean Handaxe The Swiss Army knife of the Lower Paleolithic
Although pebble tools do have a cutting edge, they are extremely simple and unwieldy. These basic tools changed, evolved, and improved over time as early hominins began to remove more and more flakes from the core of raw material, reshaping it and creating longer, straighter edges for cutting (Figure 3.29). When such a core tool assumes a distinctive teardrop shape— pointed at one end, rounded at the other, retouched to a desired size, shape, and heft—it is known as a handaxe, the signature tool of Homo erectus. The name comes originally from the French phrase coup de poing (axe wielded in the hand). But the handaxe is truly an all-purpose piece of equipment that was used for cutting, sawing, digging, bashing, and boring large holes, among other things. The handaxe is, in fact, a more complex tool than it first appears. Its final form is a shape inside a piece of stone and in the mind of the maker; a cobble must be heavily modified for the handaxe to emerge. Moreover, the handaxe is symmetrical in outline, reflecting purpose, skill, and foresight in manufacture. Handaxes are often made from small cobbles 10–15 cm (4–6 in) long. A number of much larger examples, however, also exist, some more than 30 cm (1 ft) in length. The oldest Acheulean handaxes are known from the site of Konso-Gardula in southern Ethiopia and date to 1.9 m.y.a., in association with early Homo erectus. The 700,000-year-old site of Kilombe in Kenya contains a fascinating collection of handaxes. Here, hundreds of these tools were discovered eroding out of the same geological layer. Remarkably, most of the handaxes are very similar in size and shape. Stone artifacts are made by a process of reduction, the removal of flakes from a core, and errors or mistakes
cannot be erased. Nevertheless, the symmetry and relationship between length and width of the handaxes from Kilombe is striking. Small and large implements have the same length-towidth ratio, indicating the importance of the mental image the makers had of
Figure 3.29 The evolution of the handaxe, from simple pebble-tool forms through the removal of the sides and surface of the pebble. The elongated, teardrop shape is characteristic of the handaxe, giving the artifact one pointed end and one broad end.
Concept The Acheulean Handaxe
_
99
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Man is a tool-using animal . . . without tools he is nothing, with tools he is all. —Thomas Carlyle (1795–1881)
handaxe A large, teardrop-shaped
stone tool bifacially flaked to a point at one end and a broader base at the other. reduction technique In archaeology,
a manufacturing process involving the removal (as opposed to the addition) of materials from a core that becomes the finished product. Acheulean A major archaeological
culture of the Lower Paleolithic, named after the site of St. Acheul in France. hard-hammer technique A percus-
sion technique for making stone tools by striking one stone, or core, with another stone, or hammer. soft-hammer technique A flintknapping technique that involves the use of a hammer of bone, antler, or wood, rather than stone. cleaver A tool with a broad leading
edge. burin A stone tool with right-angle
edges used for planing and engraving. Clactonian A term used for assem-
_ _
blages from the Lower Paleolithic, lacking handaxes and characterized by large flakes with heavy retouching and notches.
100
Page 100
what a handaxe should look like. The handaxes from Kilombe also document the skill the makers had in producing that image in stone. Handaxes and associated tools are referred to as Acheulean artifacts, after the original find location at St. Acheul in northern France. Floods of meltwater at the end of each glacial period downcut the rivers of western Europe, creating a series of terraces in the river valleys. On those terraces during the nineteenth century, near the towns of Abbeville and St. Acheul, prehistorians collected these signature tools of the Lower Paleolithic. Objects on the higher, older terraces were crude handaxes, with irregular edges and heavy flake scars on their surface. Acheulean handaxes from the lower, younger terraces were more symmetrical with straighter edges. A stone-on-stone method, or hard-hammer technique, was used to make the more irregular tools of the early Acheulean. A softhammer technique was used in making the younger, more regular handaxes. Mallets of bone, antler, or even wood can be used to remove flakes from stone. Lighter, soft hammers are easier to control, and the flakes that are removed are both thinner and wider. Acheulean assemblages include both handaxes and a variety of other tools, both heavy-duty pieces and smaller ones. Cleavers—handaxes with a broad, rather than pointed, leading edge—are also quite distinctive of the period. Other artifacts include a variety of flake tools such as scrapers, burins (stone tools used for gouging or engraving), and borers. Fifty years ago, Harvard archaeologist Hallam Movius described the distribution of handaxes as limited to Africa, the southern two-thirds of Europe, and western Asia. More recent research, however, has expanded the known distribution of these tools; handaxes are now known from most of Africa, Asia, and Europe. The term Clactonian refers to the nonhandaxe assemblages of the Lower Paleolithic. These assemblages represent what is called an evolved Oldo-
Chapter 3 Out of Africa: Homo erectus
wan series of artifacts, including simple pebble tools and flakes. The term is taken from the site of Clacton-on-Sea in England where this distinctive set of tools, with heavy choppers, notches, and denticulates (saw-toothed) but lacking handaxes, was found during the nineteenth century. Flake tools in these assemblages are generally blocky and irregular in shape. Clactonian assemblages lack the regularly shaped core and flake artifacts of the Acheulean. Clactonian tools appear to have been made quickly for special tasks. The distinction between handaxe and nonhandaxe assemblages is of considerable interest to prehistorians, and its significance is still not clear. Mary Leakey found both kinds of assemblages, with and without handaxes, in the same levels of Bed II at Olduvai Gorge, suggesting that the differences in the assemblages are more likely due to the activities performed than to change over time. Research in England has documented the contemporaneity of Clactonian and Acheulean assemblages, so their differences do not seem to be related to one assemblage being earlier or later than the other. Some researchers have suggested that the differences between Clactonian and Acheulean were the result of the availability of raw material, handaxes being made only where good raw material could be found nearby. Nevertheless, the presence of both kinds of assemblages at Olduvai does not support such an interpretation. The actual reason for the presence or absence of handaxes on sites of the Lower Paleolithic has not been determined.
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 101
Images and Ideas The End of the Lower Paleolithic Homo erectus in the Old World
H
omo erectus appeared shortly after the beginning of the Pleistocene, around 1.9 m.y.a. These hominins invented the handaxe and other tools; controlled fire; spread the human lineage outside Africa for the first time, to many parts of Asia and Europe; and appear to have thrived during this time. This is all the more remarkable because the Pleistocene was a very difficult and challenging period, with major changes in climate, environment, sea level, and the basic conditions of human life. It is essential to understand that during this time our ancestors began to exert their influence on their environment, to change the world around them. Clearly, the spread of early humans out of Africa was one of the most important developments in human prehistory. Following the initial appearance in southern and eastern Africa, the human species gradually increased in number and inhabited most of the more hospitable zones of the African continent. Population continued to expand, as did the geographic range of the species, and after almost 3 million years in Africa, groups of early humans began to move north toward Asia and Europe. The move out of the tropics demanded solutions to new problems in northern regions, especially the cold weather and a shortage of edible plants. Groups that had survived easily in warm climates where roots, seeds, and nuts were often available had to find new and improved ways to stay warm and obtain enough food. Although an efficient cooling system for the tropics, sweat glands were of little help to furless humans in the chilly temperate reaches of the Old World. Almost certainly, fire, shelter, and clothing—if only in the form of animal skins wrapped around the body—were used by Homo erectus in the course of their northern expansion into the continents of Asia and Europe. Fire must have been a major factor in the increasing success of human adaptation and the move into new, colder habitats. Recent evidence from the site of Gesher Benot Ya’Aqov in Israel has pushed back the earliest date for the intentional use of fire to around 800,000 years ago. Several types of wood were burned there, including willow, poplar, ash, and wild olive. Fire is used for light, warmth, protection, and cooking. Cooking with fire provides a number of advantages in addition to making food more tender and palatable. Cooking also improves the digestibility of many foods and destroys harmful toxins and microorganisms. Boiling removes juices and fats from plants and animals that are otherwise inedible. In general, the staple foods of humans are much more nutritionally rich than those of other large primates, who can subsist on leaves and fruit. Fire allows humans to make their food even more nutritional. Cooking changes the forms of starches, fats, and proteins and concentrates the nutrients in foods. High-nutrition diets were necessary for our human ancestors to meet the growing energy demands of their large brain. For Homo erectus, cooking probably made it possible to add new foods to the diet. The use of marine resources also expanded the diet of some H. erectus groups, as fish and shellfish remains from the site of Terra Amata in France document. The seas are a very rich source of food, and one that must have been exploited in the Lower Paleolithic. Images and Ideas The End of the Lower Paleolithic
< ( : < V : y < p : W 101
_
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 102
Figure 3.30 One of the remarkable, well-made spears of spruce wood from Schöningen, Germany, approximately 7 feet long and 400,000 years old.
sexual division of labor The coop-
_ _
erative relationship between the sexes in hunter-gatherer groups involving different male and female task activity.
102
Longer, colder winters in more northern latitudes also put a premium on successful predation. Meat was the primary source of sustenance during winter, when roots, nuts, leaves, and other edible plants were not available. Hunting became essential to the human way of life in colder climates. At least eight wooden spears have been found recently at the site of Schöningen in northern Germany dated to 400,000 years ago. The artifacts were found among abundant bone remains from butchered horses and several fire hearths (Figure 3.30). The spears have carved points and a length of 2.0–2.5 m (6.0–7.5 ft). The technology is impressive. The spears are made from the trunks of 30-year-old spruce trees, with the bark removed, and are sharpened at the base of the trunk, where the wood is hardest. The thickest and heaviest part of the carved shaft is about one-third of the distance from the spear point, as in a modern javelin. These weapons appear to have been designed for throwing. Such evidence indicates that the early Europeans were hunting large game by this time. It is highly unlikely that many females burdened with infants and young children could successfully hunt regularly to provide their own food. By this time, a viable relationship between males and females, incorporating food-sharing as part of pair-bonding, necessarily emerged to ensure the continuance of the human lineage. Relationships between males and females and basic family structure must have been related to these essential, adaptive changes. These connections— individual to individual, male to female, parents to offspring, kin to kin, group to group—are critical links in the chain of human society and survival. The sexual division of labor exemplifies the cooperative relationship between the sexes. Roles for males, faster and larger, as hunters and for females, with young children, as gatherers of wild plant foods emerged as an efficient, synergistic pattern for the maximization of biological capabilities. Sex, and maybe love, bonded males and females for food-sharing and reproduction; maternal instincts and extended childhoods bonded mothers to children and siblings to one another. We have no information on the precise nature of prehistoric male-female relationships—whether monogamous, several females to one male, or several males to one female. But the present universality of the human family suggests a substantial depth of time for this basic unit of society. Almost certainly, some form of protofamily emerged among Homo erectus populations if not before. Pair-bonding may have helped ensure the survival of off-
Chapter 3 Out of Africa: Homo erectus
pri31057_ch03.pp72-103.qxd
10/1/09
7:51 AM
Page 103
spring, as males began to recognize individual children as their own. The incest taboo, another human universal, may well have arisen at the same time to promote and solidify relationships beyond the immediate group. Marriage or mating outside the family ensures alliances with other families and groups, reducing the potential for conflict. Homo erectus, as a creature of the Pleistocene, was a very successful member of our lineage, expanding out of Africa, taking the human species into Asia and Europe, and setting the stage for the next step in human evolution. The evidence from the Lower Paleolithic essentially documents the appearance and spread of Homo erectus out of Africa and into Asia around 2 m.y.a. and then into Europe after 1 m.y.a. Discoveries at sites such as ’Ubeidiya and Dmanisi in Southwest Asia document the arrival of our ancestors at the crossroads of the Old World. Farther to the east, older investigations at Zhoukoudian evidence the arrival and activities of erectus groups in the east of Asia. The rigors of colder climate in much of Europe may have delayed the appearance of the first humans in this area by a million years. The first Europeans are not seen until after 1,000,000 B.P. in places such as Gran Dolina and Isernia, both in the Mediterranean part of the continent. The spears and animal prey from Schöningen boldly document our presence as big-game hunters. Back in Africa, the evidence from sites such as Kalambo Falls and Olorgesailie emphasizes our developing role as hunter-gatherers in the natural environment. Increasing brain size and more humanlike behaviors characterize the almost-2-million-year evolution of Homo erectus. Technology changed little during the Lower Paleolithic, but this time period took us from apelike to humanlike through the icy cauldron of the Pleistocene. The results of that transformation become much clearer with the arrival of Homo sapiens, described in the next chapter.
p
DISCUSSION QUESTIONS 1. What are the major characteristics of the Lower Paleolithic period? 2. What dating techniques can be used to determine the age of fossils and artifacts from Homo erectus sites? 3. What are the major characteristics of the Pleistocene epoch, and what consequences did they have for human evolution?
w w w. m h h e . c o m / p r i c e i p 6 e
For more review material and study questions, see the self-quizzes on your online learning center.
4. What kinds of evidence did the first humans outside Africa leave behind? 5. Does the handaxe represent a conceptual or a technological advance? 6. What are some of the consequences of the domestication of fire for human survival?
z
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e.
Carbonell, E., J. Castro, J. Pares, A. Perez-Gonzalez, G. Cuenca-Bescos, A. Olle, et al. (2008). The first hominin of Europe. Nature 452: 465–469. A scientific report on the most recent finds of the oldest European. Gamble, C. 1999. The Palaeolithic societies of Europe. Cambridge: Cambridge University Press. A summary of the European Paleolithic, including the first inhabitants. Klein, R. 1999. The human career, 2d ed. Chicago: University of Chicago Press. Probably the most authoritative volume on the evolution of biology and culture in the Pleistocene.
Larsen, C. S., R. M. Matter, and D. L. Cabo. 1998. Human origins: The fossil record. Longrove, IL: Waveland Press. A very readable account of human evolution. Sutliffe, A. J. 1985. On the track of Ice Age mammals. Cambridge, MA: Harvard University Press. A detailed study of the mammals of the Pleistocene. Tattersall, I., and J. H. Schwartz. 2001. Extinct humans. Boulder, CO: Westview. How we know what we think we know about human evolution.
_ Suggested Readings
103
pri31057_ch04.pp104-163.qxd
10/1/09
C
H
7:54 AM
A
Page 104
P
4
T
E
R
F
O
U
R
_ _ Figure 4.1 The Venus of Brassempouy, the small head of a woman carved in ivory, 3.65 cm high, ca. 25,000 years old, France.
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 105
The Hunters Introduction The Rise of Homo sapiens Modern humans take the stage toward the end of the Pleistocene
F
igure 4.1 shows the face of an Upper Paleolithic woman, found in France, carved in ivory and dating to more than 20,000 years ago. It is one of the very few surviving portraits of early fully modern humans in the Pleistocene. It is a remarkable statement of the interests, abilities, and directions that human beings were exhibiting at that time. Major changes in human behavior took place toward the end of the Pleistocene. For the first time, our ancestors began to exhibit behaviors that were more than just practical activities, beyond the basic necessities for survival. The genus and species Homo sapiens took the stage and became fully modern. In the Middle Paleolithic, those behaviors included burial of the dead, cannibalism, and nurturing of the weak and the elderly. By the end of the Pleistocene, our own species (Homo sapiens sapiens)—biologically indistinguishable from modern humans— had created art, invented many new tools, made tailored clothing, started counting, and spread to almost all parts of the world. The evolution of Homo sapiens is the most important development of the later Pleistocene (Figure 4.2).As we have seen, Homo erectus was the first early human form found outside Africa. By 1.2 million years ago, H. erectus fossils are known in Europe and Asia; by 100,000 years ago, H. sapiens is present. Erectus became sapiens during that time. The earlier varieties of Homo sapiens are sometimes referred to as Archaic Homo sapiens, whereas more recent Homo sapiens sapiens are termed fully modern humans (FMH). By 200,000 years ago, Europe and southwestern Asia were occupied by Homo neanderthalensis (or Homo sapiens neanderthalensis to some). Conventional wisdom today generally regards the Neanderthals as a rather specialized form that evolved from Homo erectus in the colder, more isolated areas of Europe (Figure 4.3). In the time of the Neanderthals, the Middle Paleolithic, new behaviors included burial of the dead, cannibalism, and nurturing of the weak and elderly. The earliest fully modern humans have been found in East and South Africa. The earliest known example of Homo sapiens has been found at Kibish in southwestern Ethiopia, dating to 195,000 years ago. A new project at the Cape of Good Hope at the southern tip of Africa has provided remarkable new discoveries regarding our early fully modern human ancestors. The evidence from the caves around Pinnacle Point is not in the form of fossil skeletons but rather regards the behavior of these early Homo sapiens. Several finds—small stone tools, red ochre as a pigment, the earliest known collection and consumption of shellfish—point to new kinds of food, new tools that probably required hafting, and the use of powdered mineral as a pigment or preservative. These are firsts in the
< ( : < V : y < p : W
CHAPTER OUTLINE
Introduction The Rise of Homo sapiens 105 CONCEPT The Origins of Language 109
SITE The Klasies River Mouth
Caves 111
CONCEPT Modern and Ancient
DNA 114 SITE The Valley of the
Neanderthals 117 CONCEPT The Fate of the Neanderthals 121 CONCEPT The Upper Paleolithic 123 SITE Dolni Vestonice 127
SITE The Cave of Lascaux 130 CONCEPT Portable Art 136 SITE Pincevent 138 CONCEPT Symbols and Notation 140 SITE Lake Mungo,
Australia 142 CONCEPT The Peopling of the Pacific 144 CONCEPT Radiocarbon Dating 145 SITE Beringia 147 CONCEPT Pleistocene Extinction 151 SITE Monte Verde 154 SITE Lindenmeier 157
SITE Kennewick Man 160
Images and Ideas The 105
End of the Paleolithic 161
_
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 106
Fossil sites
Figure 4.2 Human evolution in the Middle Pleistocene, 800,000–35,000 years ago. Current evidence suggests that Homo erectus expanded out of Africa into Asia sometime after 2 m.y.a. Fully modern human forms of Homo sapiens sapiens began to appear sometime after 200,000 years ago, again in Africa, moving to the Near East around 90,000 years ago and replacing the Neanderthals in Europe after 40,000 years ago.
Homo sapiens sapiens Skhul Qafzeh Cave Border Cave Klasies River Mouth
Africa
Europe Homo sapiens sapiens
Years B.P. 35,000
100,000 Homo sapiens sapiens
Homo neanderthalensis
Archaic Homo sapiens Steinheim Swanscombe Arago Verteszöllös Bilzingsleben
200,000
300,000
400,000
500,000
Homo erectus
w w w. m h h e . c o m / p r i c e i p 6 e
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
Archaic Homo sapiens
600,000
700,000
800,000
900,000
1,000,000 Homo erectus 1,100,000
1,200,000
_ _
Figure 4.3 A computer reconstruction of a young Neanderthal male. Most noticeable is the long face.
106
Chapter 4 The Hunters
archaeological record and likely point to the beginnings of the creative explosion witnessed more fully after 50,000 years ago. Beginning around 50,000 years ago, fully modern humans replaced Neanderthals in southwestern Asia and then the rest of Asia and in Europe. The mechanism for this replacement is the subject of vigorous debate among archaeologists and physical anthropologists. The major question is whether fully modern humans evolved only in Africa and spread from there across Asia and Europe, or whether they evolved in many places through the flow of genetic material between different human populations. These two competing explanations are known as the Out of Africa theory and the Multiregional theory (Figure 4.4). Different lines of evidence, from genetics and from the human fossils themselves, can be used to evaluate these theories. Genetic evidence relies on mutation rates in individuals with slightly different DNA to estimate how long ago a common ancestor existed. Such estimates suggest 200,000–140,000 years ago as the date for a common ancestor for Homo sapiens sapiens and point to Africa as the place of origin (see “Modern and Ancient DNA,” p. 114), and are close to the dates for the fossil evidence. From both genetic and fossil evidence, then, it seems very likely that fully modern humans appeared initially in Africa, sometime around 200,000 years ago. These expanding groups of modern humans created a number of remarkable innovations in human culture. Human culture, in fact, changed more during the period shortly after 50,000 years ago than it had during the previous several million years.
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 107
To the Americas between 30,000 and 15,000 B.P.
40,000 B.P.
90,000 B.P.
Initial appearance after 200,000 B.P. 40,000 B.P. (a)
(b)
Richard Klein, of Stanford University, has listed some of the innovations that mark this time: the shaping of new materials such as bone, wood, shell, and ivory into tools; the transport or exchange of raw materials, such as flint, over long distances; great diversity and specialization in artifacts; and the first art. In Europe, this period is known as the Upper Paleolithic, dating from 40,000 to 10,000 years ago. In this same general time, modern humans spread into Australia and New Guinea (around 40,000 years ago) and into North and South America (perhaps 15,000 years ago). Klein and others have suggested that these innovations represent a major change in the organization of the human brain, an advance that may be related to the emergence of complex language skills (see “The Origins of Language,” p. 109). Other changes were taking place as well. A new line of evidence concerning the use of clothing in the Paleolithic comes from a surprising source: the genetic code of lice. There are two kinds of lice in this story—head lice, which live in the hair on the head and have been around for millions of years, and body lice, which live in clothing, not on the body, and are a relatively recent species. Using mutation rates in modern lice genes, Mark Stoneking and his associates estimated that the body louse first appeared between 72,000 and 42,000 years ago. Humans must have been wearing clothing at that time to provide a habitat for this new species. As fully modern humans evolved in Africa, one assumes that their skin color was dark, pigmented by the genes for skin color as protection against the sun. As Homo sapiens migrated out of Africa into Asia and Europe, clothing would have become more important as protection against the rigors of Pleistocene winters. A recent study of a gene skin color suggests that the appearance of pale skin in European populations may be quite recent, perhaps only in the past 10,000 years. This chapter tells the story of the Neanderthals and of the appearance, expansion, and spread of Homo sapiens to virtually all parts of the world, as evidenced at sites such as Klasies River Mouth in South Africa, the Neander Valley in Germany, Shanidar in northern Iraq, Dolni Vestonice in the Czech Republic, and Lascaux and Pincevent in France (Figure 4.5). The evidence from Lake Mungo in Australia and from Monte Verde, Lindenmeier, and Kennewick in the Americas documents the movement of humans to Australia and the New World during the past 40,000 years. (A useful distinction is often made between the Old World—Africa, Asia, and Europe—and the New World of North and South America. The New World is conventionally considered new because it was discovered by Europeans at a relatively late date.) The Pleistocene and the Paleolithic came to an end some 10,000 years ago. The glaciers retreated as warmer temperatures prevailed, and our present epoch began. The Paleolithic closed as the human species on six continents began to adapt to the warmer conditions of the post-Pleistocene, or Postglacial, as it is also known.
Figure 4.4 The two major competing theories about the evolution of fully modern Homo sapiens. (a) The Out of Africa theory involves the evolution of H. sapiens from H. erectus in Africa and the spread of that species. (b) The Multiregional theory argues that H. sapiens evolved from H. erectus in several places and that interbreeding kept them similar. Most of the archaeological and genetic evidence supports the Out of Africa theory.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the controversy surrounding the emergence of our species, see the Internet exercises on your online learning center.
Introduction The Rise of Homo sapiens
107
_
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 108
Hebior/ Schaefer
Kennewick
Pincevent Lascaux ATLANTIC OCEAN
Lindenmeier
Ushki
Neander Valley Dolni Vestonice Cosquer
PACIFIC OCEAN Kibish Equator
PACIFIC OCEAN INDIAN OCEAN Pinnacle Point
Sites discussed in the chapter Monte Verde
Lake Mungo
Klasies River Mouth
Geographic distribution of human population by the end of the Pleistocene
Kibish Pinnacle Point Klasies River Mouth Neander Valley Dolni Vestonice Lascaux Cosquer
Figure 4.5 Location of and timeline for Homo sapiens and the Neanderthals in the late Pleistocene and the Paleolithic.
Pincevent Lake Mungo Ushki Monte Verde Hebior/Schaefer Lindenmeier Kennewick
_ _
200
175
150
125
100 Thousands of years ago
108
Chapter 4 The Hunters
75
50
25
0
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 109
Concept
<
The Origins of Language Why we first spoke may be more important than when
The origin of human speech and language is one of the most fascinating aspects of human adaptation and evolution, yet perhaps the most difficult to explain. Modern languages contain hundreds of thousands of words. We use words to convey information about every aspect of our lives. Shakespeare’s vocabulary is estimated to have been 24,000 words; a newspaper reporter uses approximately 6000 words; the average person on the street has a speaking vocabulary of some 3000 words. We have a natural interest in when and where this ability to communicate with the spoken word originated. Language did not appear suddenly at some point in the past, without antecedents; it evolved gradually from the utterances and cries of early primates to its modern forms. On the one hand, the English language today is a huge complex of vocabulary, grammar, and structure, to which many new words are added each year. On the other hand, most animals make sounds. Monkeys vocalize to express emotion but do not have voluntary control over vocalization; for one thing, they lack the vocal apparatus humans have (Figure 4.6). Chimpanzees, however, have a repertoire of 20 or more vocalizations and gestures for expressing their needs. Although these apes can manipulate symbols, they are unable to connect more than two or three concepts in a single phrase. To understand the evolution of language from gestures and cries to its complexity today, we must appreciate the path of its development. Studies of the physical remains of early humans provide substantial information about language use by early hominins. Discovery of the hyoid bone in a Neanderthal burial from Kebara Cave in Israel showed that it was no different from our own. The hyoid bone holds the muscles of the tongue
Soft palate Epiglottis
Soft palate Pharynx
Pharynx Larynx Hyoid bone
to the throat. The similarity between the hyoid bone in Neanderthal and in modern-looking Homo sapiens suggests that speaking abilities would also have been similar. Endocasts of the inside of the skull provide the only direct evidence of brain organization (Figure 4.7). The cerebrum, the upper portion of the brain, is primarily concerned with the complexities of behavior. This area is large and developed in higher primates. The size of the cerebrum, its convoluted surface, and the extent of wrinkling have increased over the course of evolution of the human species from our primate ancestors. The organization of the cerebrum is critically important. In modern humans, the front of the brain is much larger than the back, and the sides of the brain are well developed, in contrast to the brains of chimpanzees and other apes. The two sides of our brain operate cooperatively to direct and control different aspects of our behavior and activities. This division in the organization and operation of the brain is called lateralization. One side of the brain controls language, and the other side regulates motor skills and perception. Lateralization is essential for language, because the processing of word
Epiglottis Larynx
Figure 4.6 The vocal apparatus of a chimpanzee (left) compared with that of a modern human (right). Notice the more complex vocal cords in humans. The human pharynx is more curved, is deeper in the throat, and produces a great variety of sounds.
hyoid bone A delicate bone in the neck that anchors the tongue muscles in the throat. endocast A copy or cast of the inside of a skull, reflecting the general shape and arrangement of the brain and its various parts. lateralization The division of the
human brain into two halves.
Concept The Origins of Language
_
109
pri31057_ch04.pp104-163.qxd
10/1/09
Figure 4.7 Endocasts from South African australopithecines. Notice the details of the blood vessels and other features of the brain that have been preserved in the fossil casts.
_ _
110
Chapter 4 The Hunters
7:54 AM
Page 110
strings must occur in close proximity in the nerve cells of the brain. Individuals with speech problems are probably sequencing words and controlling speech from both sides of the brain. Studies by Ralph Holloway, of Columbia University, have shown that the pattern of lateralization in fossil endocasts goes back well into the Pleistocene and probably to australopithecines as well. Dean Falk, of the State University of New York, has also been involved in the study of endocasts, pointing out that Broca’s area, a region of the brain involved in the control of language, is larger in Homo ha-
bilis than in their contemporaries, the australopithecines from East Africa. Research on human use of language has recently turned to the hypoglossal nerve, which controls the movement of the tongue. This nerve is twice as thick in humans as in chimpanzees. This thick nerve began to appear around 500,000 years ago with larger-brained members of the genus Homo, and it may be at that point that more sophisticated language became possible. Other fossil evidence indicates that the necessary mouth and throat anatomy for a spoken language was not in place until 150,000 years ago. It is still not clear when language emerged or whether the process was gradual or sudden. Certainly, many of the activities of our early Pleistocene ancestors would have required some form of communication. Food-sharing, social organization, and other distinctly human characteristics imply a system of verbal expression. These abilities must have evolved and expanded through time, as both brainpower permitted and need required. One of the major unresolved issues in the development of language is the shift from a primitive language, like that of small children, to a syntactic one, with grammatical rules and structure. This development may be related in part to further changes in the human brain. Some linguists suggest that all the world’s languages evolved from a common “mother tongue,” and a few would even suggest a date of 100,000 years ago for this common language. Needless to say, this is highly speculative, but it does suggest that future research may provide more information on the development of human languages. One of the more striking developments in human prehistory was the “creative explosion” that occurred about 50,000 years ago, around the beginning of the Upper Paleolithic. The changes witnessed in this period may well be related to significant advances in our language abilities.
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 111
Site
:
The Klasies River Mouth Caves One of the longest continuous sequences of human habitation in the world
A series of caves cluster at the mouth of the Klasies (CLASS-ease) River where it empties into the Indian Ocean in South Africa (Figures 4.8 and 4.9). The caves were originally cut by wave action against the high sandstone and shale cliffs in this area at a time when sea level was higher than it is today. The caves and the sandy area in front of them were a hospitable place for human residence for more than 60,000 years, from 120,000 to about 60,000 years ago. Occupation remains were so abundant here that the accumulated debris in one of the caves had completely buried the opening of a lower cave. The attractions of the site for repeated residence included the shelter of the caves; the moderate climate; the immediate availability of marine foods such as shellfish, seals, and even beached whales; nearby fresh water; access to large and small mammals living along the river; and good-quality stone for toolmaking. The 60,000 years of deposits are 20 m (65 ft) deep (the height of a 6-story building) and span the entire Middle Stone Age (MSA) of southern Africa (Table 4.1). (In Africa, the terms
Early, Middle, and Late Stone Age are used to distinguish the archaeological divisions of the Paleolithic.) Although the deposits at the Klasies River Mouth are enormously deep, they accumulated at a rate of only 5–10 cm (2–4 in) every 100 years.
Figure 4.8 Caves 3 and 4 at Klasies River Mouth, along the southern coast of South Africa.
Figure 4.9 The Klasies River Mouth area coastline, showing the location of caves, artifacts, and resources. Klasies Klasies River River
Artifacts in breccia Cave 2
Sea cave
Cave 1
Cave 3 Cave 4 Cave 5 N
Sea cave 6–8 m beach
Rock shelf exposed at low tide with mussels, winkles, alikruikel, oysters, etc.
INDIAN OCEAN 0
500 m
_ Site The Klasies River Mouth Caves
111
Page 112
Deposits
Environment/Resources
10,000 B.P.
Later Stone Age
Seacoast 60 km distant; some limpets
70,000 B.P.
No archaeology Middle Stone Age III
Sand with thin layers of silt and clay Shell middens and sand
112
Chapter 4 The Hunters
wildebeest
MSA: phase I
oribi
MSA: phase IIa
grysbok
_ _
vaalribbok
MSA: phase III
bastard hartebeest
MSA: phase IV
hartebeest
Dating these deposits is difficult. The lower layers are beyond the range of radiocarbon dating. Other dating techniques have been employed, including oxygen isotope ratios (see “Climate and Environment in the Pleistocene,” Chapter 3, p. 81). Geological cores removed from deep ocean sediments contain microscopic shells that can be measured for water temperature. It is thus possible to construct graphs of changing water temperature extending hundreds of thousands of years into the past. It is also possible to measure the oxygen isotope ratio in shells found in archaeological sites. The shells in stratigraphic layers from the Klasies River Mouth (KRM) Caves were measured and the results compared with the ocean sediment curve for which the age was already known. This correlation and other dating methods—including electron spin resonance, uranium disequilibrium, and amino acid dating—suggest that the age of the MSA deposits extends from 120,000 to 60,000 years.
Figure 4.10 Changing utilization of species over time at Klasies River Mouth documents the variety of species that were hunted by the inhabitants.
MSA: phase IIb
Cooler; increased forest and bush; marine species: seals, birds, dolphins, whales, limpets High sea level; temperate; mixed forest/ grassland; seals, dolphins, penguins, shellfish (limpets and mussels)
blue antelope
Middle Stone Age I
mountain reedbuck
120,000 B.P.
Shell middens and sand Shell middens and sand; rubble and artifacts in cave Shell middens and sand
southern reedbuck
Middle Stone Age IIb Middle Stone Age IIa
Cooler and wetter, drop in sea level; shellfish (mussels, turbot); open grasslands
Excavations over a number of years by Ronald Singer and John Wymer, of the University of Chicago, and Hilary Deacon, of the University of Stellenbosch in South Africa, have exposed the buried occupation layers and revealed a number of pieces of evidence of major importance for understanding Old World prehistory. Fully modern humans (Homo sapiens sapiens) appeared here earlier than anywhere else in the world. A handful of fossil fragments from KRM and other sites in South Africa indicate that modern-looking humans were present in this area by 100,000 years ago. The human remains from Klasies are fragmentary and show breakage, cutmarks, and burning that suggest cannibalism. Evidence from animal bones, studied by Richard Klein of Stanford University, documents a successful economy throughout the Middle Stone Age. A wide range of animal species is represented, and both large and small mammals were abundant (Figure 4.10). Porcupine, grysbok (a small antelope), eland (a large antelope), giant buffalo, rock hyrax (a small mammal), and the Cape fur seal were the most common. The site also records the early use of marine foods, evidenced by limpet and mollusk shells, and the bones of penguins and seals. The shells could have been collected all along the coast by wading at low tide. There are abundant carbonized organic remains in the deposits, and Deacon believes that these may represent roots and tubers that were collected and eaten by the inhabitants. The faunal evidence from KRM can also be used to examine the hunting abilities of these Stone Age people in South Africa. Consider two possible patterns of death in a population of animals: (1) attritional, in which death is by natural causes, such as predation,
springbok
Period
giant buffalo
Age
Cape buffalo
TABLE 4.1 Sequence of Layers and Corresponding Changes in the Environment at Klasies River Mouth
eland
7:54 AM
kudu
10/1/09
bushbuck
pri31057_ch04.pp104-163.qxd
pri31057_ch04.pp104-163.qxd
Nelson Bay Cave, Late Stone Age (present interglacial deposits) Klasies River Mouth, Middle Stone Age (last interglacial deposits)
10/1/09
7:54 AM
Page 113
Eland
Cape buffalo
Bushpig
6
39
38
80
29
3
disease, accident, and old age; and (2) catastrophic, in which a natural disaster, such as flood, epidemic, or mass hunting, simultaneously kills most of the members of the population. The catastrophic pattern would provide an almost complete picture of the ages and sexes of the living population, whereas an attritional pattern would be dominated by young and old animals, those more susceptible to predation and disease. Prime-age adults would be largely missing in the attritional pattern. Comparison of the remains of two species of animals from KRM indicates that the eland, a large antelope, is represented by the catastrophic pattern, whereas the pattern for Cape buffalo is attritional. The eland is a relatively docile animal that can be driven into traps or falls by hunters, which probably explains the catastrophic death pattern. Cape buffalo are much more recalcitrant and dangerous. The high proportion of young in the pattern for the Cape buffalo was likely the result of selective hunting. Thus, animals that could be driven are represented by catastrophic death patterns, whereas more aggressive species were probably hunted individually with weapons. The higher number of elands at KRM, even though this species was much less common than the Cape buffalo in the environment, suggests that the MSA people were not particularly good hunters and that drives may have been more effective than stalking. If we compare the remains of these two species from a younger site at nearby Nelson Bay Cave, dating to perhaps 10,000 years ago, we see that the death pattern for the buffalo is the same and that eland remains are very rare (Figure 4.11). This pattern suggests that
Figure 4.11 The minimum number of eland, Cape buffalo, and bushpig at the sites of Klasies River Mouth and Nelson Bay Cave. Buffalo and bushpig are more difficult to hunt, and their abundance in the Holocene deposits of Nelson Bay Cave likely reflects the better technology of the more recent hunters.
the later groups were better hunters than their counterparts at the KRM caves. The argument is supported by evidence for bows and arrows at the time of the Nelson Bay Cave occupation. On the other hand, Deacon believes that people of the Middle and the Late Stone Age were behaviorally similar, hunting smaller animals and both hunting and scavenging larger ones. Artifacts, hearths, bones, and shells are found in the oldest layers right next to the human skeletal remains dating to approximately 120,000 years ago. Shellfish, mostly brown mussel, were collected in quantity, and shell middens accumulated. One of the shell middens is approximately 5 m (16 ft) high and is as extensive as any Late Stone Age midden. This evidence is important because it shows regular shellfish collecting by the inhabitants and also systematic disposal of food refuse in localized heaps. This pattern suggests that people were living by the same rules for the use of space and cleanliness in both the Middle and the Late Stone Age. Deacon points out that the large mammal bones and shell middens are the most obvious food remains. Remains of plant foods appear as carbonized materials around the hearths and indicate that plant gathering was also an important activity. It is this carbonized material that creates distinctive black horizons that are common features at the site. Deacon reports similar patterns of carbonized vegetation around the hearths of more recent inhabitants of South Africa. Deacon’s perspective suggests that the general way of life in this area may not have changed from 100,000 years ago until the end of the Pleistocene. Site The Klasies River Mouth Caves
_
113
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 114
Concept
<
Modern and Ancient DNA Archaeology and genes
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on how the study of genetics impacts our understanding of modern human evolution, see the Internet exercises on your online learning center.
DNA (deoxyribonucleic acid) is the genetic material of all life. DNA is a long molecular chain of units called nucleotides, each composed of one of four base units (adenine, cytosine, guanine, or thymine). There are some 50,000 genes in the DNA of a human being. Genetic information is coded in a sequence of nucleotides in individual genes. Individual genes that determine the growth and characteristics of individuals are segments of DNA molecules. Molecules of DNA also have “intergenic” spaces between genes with no biological information that are important as markers for differences between individuals and between populations. Two major types of studies, focused on either modern or ancient DNA, are being done to investigate the human past. Modern populations are studied to identify genetic differences and the
0
Figure 4.12 Relative distances among various modern species as determined by nucleotide substitutions. The number of substitutions is shown at the bottom of the tree graph, and the estimated age of the divergence between two species and their common ancestor is shown at the top. For example, cows and dogs probably separated from a common ancestor around 70 m.y.a.
_ _
114
Chapter 4 The Hunters
20
40
time at which groups of people diverged in the past. These studies generally use DNA in blood or other cells. DNA studies of modern populations have been used to estimate the time of the first appearance of modernlooking humans and to examine the spread of Neolithic farmers into Europe, for example. Evidence from genetic studies also provides some understanding of the point at which different animal species separated—for example, when humans became a species distinct from an apelike ancestor. Biological scientists have developed a “molecular clock” to estimate the number of years since different modern species separated from a common ancestor (Figure 4.12). The mechanism for this clock is the changes over time in the amino acids that make up DNA. Known as nucleotide substitutions, these changes are observed as
Millions of years ago 60 80
100
120
Man Chimpanzee Gorilla Gibbon Rhesus Aethiops Rabbit Rodent Dog Horse Donkey Pig Llama Sheep Goat Cow Kangaroo 0
25
50 Nucleotide substitutions
75
100
pri31057_ch04.pp104-163.qxd
Mitochondrion
10/1/09
7:54 AM
Page 115
Ribosomes
Chromosome
Nucleus
Cytoplasm
The chromosome consists of DNA wound around binding proteins
Chromosome unwinding
mismatches in the genetic material of two species. The number of mismatches in the chromosomes of two species correlates closely with the evolutionary distance between the species and the amount of time they have been separate. Our closest relatives are the chimpanzee and the gorilla. Comparison of amino acid sequences in humans, chimps, and gorillas indicates that the three species diverged between 5 and 4 m.y.a.—a date very close to that for the earliest fossil hominins. There is also intriguing modern genetic evidence regarding the evolution of Homo sapiens sapiens. This evidence comes from studies of mitochondrial DNA (mtDNA), genetic material that is assumed to mutate at a relatively rapid and constant rate (Figure 4.13). Because this type of DNA is inherited directly through the maternal line, it provides a continuous trail back into the past. Graphing mutations in mtDNA results in a tree with the oldest changes as the trunk and more recent mutations as the branches. The number of mutations separating two
individuals should be a function of how far back in time they shared a common maternal ancestor. Analysis of this DNA from a number of women from around the world allows a map to be drawn of the spread of modern Homo sapiens after 200,000 years ago (Figure 4.14). On the basis of the number of accumulated mutations in the mtDNA, researchers concluded that Homo sapiens first appeared in southern Africa between 170,000 and 130,000 years ago. This is known as the “African Eve” hypothesis. Similar studies of the Y chromosome, present only in males, have indicated a similar date of divergence, around 200,000 years ago. These studies also suggest that modern humans moved out of Africa around 130,000 years ago, into Asia after 70,000 years ago, and into Europe after 50,000 years ago. Modern humans arrived in Australia after 40,000 years ago and came to the New World after 30,000 years ago. The second kind of study involves ancient DNA, genetic material in the nucleus of cells extracted from the remains of ancient plants and animals.
Figure 4.13 DNA is found both in the nucleus (central core) of cells and in floating structures within the cell known as mitochondria.
mitochondrial DNA Genetic material in the mitochondria of human cells that mutates at a relatively constant rate. ancient DNA Genetic material preserved in archaeological remains of bones and plants that can be studied for information about past genetic relationships.
_ Concept Modern and Ancient DNA
115
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 116
New World population Origin: 15,000 to 30,000 years ago
European population Origin: 40,000 to 50,000 years ago
Asian population Origin: 50,000 to 70,000 years ago
Northeast African group
African populations Origin: 150,000 to 200,000 years ago Population: 23,000 to 450,000
Immigration from Africa About 137,000 years ago 200 to 500 or more individuals
0
0
2500
2500
Figure 4.14 Analysis of modern mitochondrial DNA provides a map of the origin and movement of modern humans from Africa across the continents of the Old and the New World.
_ _
116
Chapter 4 The Hunters
Australo-Melanesian population Origin: 40,000 to 60,000 years ago
5000 Miles
5000 Kilometers
Discovery of DNA preserved in prehistoric human bone was first reported in 1989. Since that time, numerous studies have looked for and found DNA in ancient materials. Samples from human bone, for example, can provide information on the sex or genealogy of an individual or on genetic relationships between populations and migration. In many cases, molecules in ancient DNA have been damaged by decay and degradation over time and are found often as short segments of the larger molecule. This breakdown of the molecule makes it more difficult, but not impossible, to reconstruct original genetic information. Analysis of ancient DNA has been made possible by the development of the technique known as the polymerase chain reaction (PCR), which results in the cloning of large quantities of material for analysis even when only a very small sample is available—often the case with archaeological remains. Theoretically, even a single molecule, as well as badly degraded segments of molecules, can be analyzed with the help of the PCR technique.
Contamination is a significant problem. Since only small amounts of ancient DNA are present in samples, any contamination from living humans during excavation or laboratory analysis can mask or hide the prehistoric materials. Researchers must use great caution when removing and preparing samples for such analyses, and most researchers insist on duplication of their results in another laboratory.
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 117
Site
:
The Valley of the Neanderthals Close relatives with strange habits
In 1856, 3 years before Charles Darwin published his extraordinary treatise On the Origin of Species, proposing natural selection as a mechanism for evolution, pieces of an unusual skeleton were unearthed in a limestone cave in the valley of the Neander River, near Düsseldorf, Germany. Before this discovery, there had been no acceptance of human forms earlier than Homo sapiens and only limited awareness of a concept such as human evolution. Leading authorities first described the bones from the Neander Valley as those of a deceased Prussian soldier, a victim of Noah’s flood, or a congenital idiot—but definitely not an early human ancestor. Gradually, however, more examples of these individuals came to light. In 1886, at the cave of Spy (pronounced “spee”) in Belgium, two similar skeletons were discovered in association with early stone tools and the bones of extinct animals, clearly proving the antiquity of humans in Europe. In 1913, French physical anthropologist Marcellin Boule published a study of an arthritic Neanderthal skeleton from the site of La Chapelle-auxSaints (Figure 4.15). In this report, he described the finds from Europe as a new species, designated Homo neanderthalensis. Unfortunately, Boule did not acknowledge the discoveries of Homo erectus from Java and saw the Neanderthals as somewhere between ape and human. In Boule’s own words (translated from the French), “The brutish appearance of this muscular and clumsy body, and of the heavyjawed skull, declares the predominance of a purely vegetative or bestial kind over the functions of the mind” (1913). His work resulted in a view of Neanderthals as slow in wit, gait, and habit—an idea that continues in some quarters even today.
Gradually, however, as more Homo erectus and australopithecine specimens were reported and accepted into the family tree, Neanderthals came to be recognized as closer to modern humans. Today, they are usually classified as a member of our own genus but are distinguished at the species level, as Homo neanderthalensis.
Figure 4.15 A Neanderthal skeleton from La Chapelle-aux-Saints, France, discovered early in the twentieth century. The femur and the vertebrae are deformed by arthritis. These remains led the scientist Marcellin Boule to describe Neanderthals as brutish and slow in wit, gait, and habit.
_ Site The Valley of the Neanderthals
117
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 118
TABLE 4.2 Major Characteristics of Neanderthals Compared with Modern Humans
_ _
Figure 4.16 A reconstruction of the soft tissue on a Neanderthal skeleton. Forensic scientists have studied the attachment of muscle, fat, and skin to the bones of the face and skull and are able to reconstruct a likeness of a deceased individual. Such skills have been applied to a Neanderthal, with the result shown here. Body hair was intentionally not added to the reconstruction, with the exception of the eyebrows.
118
Chapter 4 The Hunters
Trait
Homo neanderthalensis
Homo sapiens sapiens
Forehead Brow ridges Face Cranial capacity Protrusion on back of skull Chin Appearance of skeleton
Sloping Moderate Slightly forward 1450 cc Present Absent Robust
Vertical Absent Below forehead 1400 cc Absent Present Gracile
Neanderthals were short and stocky, averaging about 1.5 m (5 ft) in height, with bowed limbs and large joints supporting a powerful physique (Figure 4.16; see also Figure 4.21). Fossil skeletons of Neanderthals are recognized today by several distinctive features in the skull and teeth (Table 4.2). The cranium is relatively low, and the face is long. Prominent brow ridges—bony protrusions above the eyes—and generally heavy bone structure give the skull a distinctive look (Figure 4.17). The face is large, the forehead slopes sharply backward, and the nose and the teeth sit farther forward than in any other hominin, giving the entire face an elongated appearance. This face is probably the result of a combination of factors, including adaptation to the cold. The average brain size of the Neanderthals is slightly larger than that of modern humans, probably a consequence of their heavier bone structure. A distinctive shelf or protrusion at the back of the Neanderthal skull is known as an occipital bun. The front teeth are often heavily worn, even the deciduous teeth of young children, suggesting that they were used for grasping or heavy chewing. Intriguing small scratches often occur on the front teeth, usually running diagonally. These marks are thought to be the result of “stuff-and-cut” eating habits, in which a piece of meat was grasped in the teeth and a stone knife was used to cut off a bite-size piece at the lips. Occasionally the knife must have slipped and scratched the enamel of the front teeth. Most of the scratches run
from upper right to lower left, although about 10% are in the opposite direction. Such evidence confirms that right- or left-handedness among humans was common by this time. The skeleton of the Neanderthals differs somewhat from that of fully modern forms, although they had the same posture, dexterity, and mobility. Neanderthal bones are generally described as robust; they had heavier limb bones than fully modern humans, suggesting much greater muscular strength and a more powerful grip. This strength is also evident in the shoulder blades and neck, and on the back of the skull, where heavy muscle attachments are noticeable. Shoulder blade muscles would have provided the Neanderthals with strong, controlled downward movements for making stone tools or thrusting spears. The robust appearance of the Neanderthals may be related to the strength and endurance required for longdistance travel over irregular terrain or to climate. Study of the Neanderthal body indicates that their stout shape is similar to that of the Eskimo, perhaps reflecting an adaptation to cold temperatures. Or perhaps Neanderthals had to be stronger to accomplish physically what fully modern humans accomplished with sophisticated tools. Neanderthal skeletons exhibit more traumatic damage, especially to the head and neck, from accident or violence than many modern populations, perhaps from close encounters with large game. The Neanderthals lived to their late thirties or mid-forties, a rather long life span in antiquity.
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 119
Classic Neanderthal Classic Neanderthal
Early Upper Paleolithic modern human
Recent man from Faeroe Islands
Early Upper Paleolithic modern human Recent man from Faeroe Islands
Neanderthal populations are generally associated with the manufacture of a variety of flake tools in groups of artifacts termed Mousterian (mooseTEER-e-an) assemblages, after the site of Le Moustier in France. The Mousterian belongs to the Middle Paleolithic, dating to approximately 120,000–40,000 years ago in Europe. Although handaxes continued to be made, large retouched flakes and Levallois pieces, from a technique for making thin flakes with a lot of cutting edge, are the major hallmarks of the period. Flakes were shaped into a variety of tools for more special purposes (Figure 4.18). Neanderthal fossils and/or Mousterian assemblages are found primarily in Europe and southwestern Asia (Figure 4.19). The Neanderthals were largegame hunters. Isotopic studies of Neanderthal bones document a carnivorous diet. Their prey varied across Europe; reindeer were hunted primarily in the west, and mammoths were hunted in eastern regions. Neanderthals were apparently an indigenous adaptation in Europe, well adapted to the cold conditions of the Pleistocene. However, an extremely cold period around 75,000 years ago may have pushed some Neanderthal populations southeast into Southwest Asia and eastward into western Asia (see “The Fate of the Neanderthals,” p. 121). The easternmost Mousterian sites are known from the Altai Mountains, around Lake Baikal and into western Mongolia. Cultural innovations during this period include the first intentional
Figure 4.17 Differences between the skulls of a Neanderthal (La Ferrassie A), an early fully modern individual from the Upper Paleolithic (Predmosti 3), and a recently deceased male Homo sapiens sapiens from northern Europe. The bulging forehead, the presence of a chin, reduced brow ridges, and the absence of a large protrusion on the back of the skull, known as an occipital bun, characterize modern humans. (From Athena Review, Vol. 3, No. 2, p. 54, 2002. Image courtesy of Athena Review and C. Loring Brace.)
The fact that Neanderthals buried their dead is surprising to many people. What do you think this practice may have meant?
y
The past is a foreign country; they do things differently there. —L. P. Hartley
brow ridge That part of the skull above the eye orbits. occipital bun A distinctive shelf or
protrusion at the base of the skull; a feature usually associated with Neanderthals. robust “Big-boned,” heavy, thick-
walled skeletal tissue. Robust early hominins had very large teeth. Mousterian A term describing the
stone tool assemblages of the Neanderthals during the Middle Paleolithic, named after the site of Le Moustier in France.
Figure 4.18 Typical heavy flake tools from the European Mousterian.
Levallois A technique for manufacturing large, thin flakes or points from a carefully prepared core.
Site The Valley of the Neanderthals
119
_
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Figure 4.19 The distribution of Homo neanderthalensis and Mousterian sites in Europe and the Near East. Homo sapiens sapiens were living in most of the rest of Asia and Africa during this time.
Page 120
EUROPE
CHINA
PACIFIC OCEAN AFRICA JAVA INDIAN OCEAN
Neanderthals Non-Neanderthals and premodern humans
Figure 4.20 An artist’s speculative reconstruction of the Neanderthal burial ground at La Ferrassie, France. The graves of several infants and adults were uncovered here during the 1800s.
_ _
Figure 4.21 An artist’s reconstruction of a Neanderthal female.
120
Chapter 4 The Hunters
burial of the dead in graves (Figure 4.20), sometimes accompanied by flowers, tools, or food. The presence of these materials in graves certainly implies concepts of death as sleep or of life after death. More exotic practices emerge as well, difficult to understand or explain from our modern perspective. Several examples of broken and burned human bones have been found among the remains of other animals in deposits belonging to the Middle Paleolithic period. At the cave of Krapina in Croatia, the bones of at least 13 human individuals were found, along with those of various herbivores and other animals. The human bones had been burned, split to extract marrow, and
treated like the bones of the animals that had provided meals for the occupants of this site. At the Grotte de l’Hortus in southwestern France, similar evidence of cannibalism was found. Heavily fragmented Neanderthal bones from at least 20 people were scattered among the numerous bones of small wild goats. Most of the human bones are skull and jaw fragments, and many of the individuals were over age 50. Whether such practices were rituals of consecration of the dead, or the bones were simply individuals from enemy groups that were added to the larder, is not known. It has also been argued that the bones may have been accidentally burned and broken by later inhabitants at the site.
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 121
Concept
<
The Fate of the Neanderthals A peaceful or violent end?
Between approximately 45,000 and 25,000 years ago, Neanderthals became extinct and were replaced by fully modern humans in Europe and western Asia. The fate of the Neanderthals is open to question: Were they completely replaced by fully modern humans, perhaps violently, or did they interbreed and simply disappear in the mix? The evidence on this transition is quite different in Southwest Asia and in Europe. Current evidence from newly dated sites in Southwest Asia suggests that the first fully modern humans appeared in this area as much as 100,000 years ago (Figure 4.22). At Qafzeh Cave and the site of Skhul in Israel (Figure 4.23), and elsewhere in Southwest Asia, the bones of Homo sapiens sapiens are found in layers with Mousterian tools, dating to 90,000 years ago. At other sites, such as Shanidar and Kebara, Neanderthal skeletons have been found dating to between 75,000 and 45,000 years ago. It is entirely possible that the Neanderthals found in Southwest Asia moved there from Europe during a period of intense cold. It appears that fully modern humans coexisted with Neanderthals in Southwest Asia until around 45,000 years ago. The earliest evidence for Upper Paleolithic technology comes from East Africa more than 50,000 years ago and includes bone tools and pendants, along with a standardized set of artifacts. In Europe, the transition is less clear, and evidence for the first fully modern humans shows them appearing later. Neanderthals are first known in Europe by approximately 250,000 years ago, yet the earliest bones of fully modern humans do not appear in this area until after 40,000 years ago. Recent evidence from new, calibrated AMS radiocarbon dates (see “Radiocarbon Dating,” p. 145) has
Years
Oxygen isotope stage
B.P.
20,000 30,000 40,000
Qafzeh
3
Upper Paleolithic
Modern 50,000 Middle Paleolithic
Kebara 60,000 Neanderthal 70,000
4
Shanidar
80,000
Neanderthal
90,000 100,000
5
Qafzeh Skhul
Modern
110,000
shown that the expansion of fully modern humans was faster than previously thought and that the period of their coexistence with Neanderthals was considerably shorter. Thus, Homo sapiens sapiens arrived in southeastern Europe from the Near East around 46,000 years ago and reached western Europe within 4000–5000 years. The Neanderthals may have been largely gone from Europe by 40,000 years ago, perhaps coinciding with the onset of one of the coldest periods of the Pleistocene, as well as the arrival of fully modern humans. Although there appear to have been some surviving populations of Neanderthals in some parts of Europe, for the most part
Figure 4.22 The sequence of hominins from major sites in the Near East. This important evidence indicates that fully modern humans were present in this area by 100,000 years ago, before the Neanderthals. Neanderthals were probably forced out of Europe and into western Asia during an extreme period of cold. The Neanderthals in the Near Eastern sites date to between 70,000 and 45,000 years ago. Technological changes from Middle to Upper Paleolithic took place around 50,000 years ago and were not related to the human subspecies present. (Reprinted by permission of Waveland Press, Inc. from Larsen et al., Human Origins: The Fossil Record, 3/e. [Long Grove, IL: Waveland Press, Inc., 1998]. All rights reserved.)
Concept The Fate of the Neanderthals
_
121
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 122
Figure 4.23 A view of the caves at Mount Carmel, Israel. The caves of Tabun and el-Wad are in the center of the photo. The site of Skhul is out of view to the left. Many of the Paleolithic sites in the Near East are found in similar caves that formed in the limestone geology of the region.
Homo sapiens replaced Homo neanderthalensis very quickly. The question of the fate of the Neanderthals remains unsolved. Why did they disappear? Several possibilities have been suggested in both scientific and popular literature. Were Neanderthals simply conquered and slain by advancing groups of technologically superior Homo sapiens sapiens? Was it a major period of cold climate in Europe? Two recent discoveries of human skeletal remains suggest some interbreeding of Neanderthals. The oldest fully modern human skull in Europe, dating to 35,000 years ago at the cave of Oase in Romania, combines characteristics of modern humans and
Chimpanzees
Neanderthals
_ _
Figure 4.24 A dendrogram of the genetic relationship between modern humans, Neanderthals, and chimpanzees, based on the study of ancient DNA in several European Neanderthals.
122
Modern Humans
4 million– 6 million 550,000– years ago 690,000
Chapter 4 The Hunters
120,000– 150,000
Neanderthals. A child burial at the cave of Lagar Velho in Portugal, dating to 24,500 years ago, has been described as the product of Neanderthal and modern mating. The long span of time between these two dates also suggests a very long period of contact between the two subspecies. On the other hand, there is genetic evidence to the contrary. Did Neanderthals disappear into the gene pool of modern-looking humans as smaller numbers of Neanderthals interbred with larger numbers of Homo sapiens sapiens? Ancient DNA extracted from a newly excavated 40,000-year-old Neanderthal bone from Feldhofer Cave, Germany, and DNA from bones at several other places in Europe and western Asia suggests that there was little genetic relationship, and thus no mating, between the Neanderthals and the fully modern humans who replaced them (Figure 4.24). From this evidence, the end of the Neanderthals may have been more violent than romantic, or perhaps a combination of cold conditions and aggressive new neighbors. The jury is still out in this scientific trial.
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 123
Concept
<
The Upper Paleolithic The arrival of Homo sapiens sapiens Magdalenian Solutrean Perigordian
HOLOCENE
Warmer
Aurignacian
Cooler
Temperature change
Châtelperonnian
40
38
36
34
32
30
28
26
24
22
20
18
16
14
12
10
Thousands of years ago
Figure 4.25 A chart of the chronology, climatic changes, major cultural periods, and typical artifacts of the Upper Paleolithic. The differences between these European cultures were much greater during the Upper Paleolithic than during the Middle Paleolithic.
The Upper Paleolithic is characterized by a variety of innovations that developed over the last 40,000 years or so of the Pleistocene. These include the arrival of anatomically modern humans in Europe; the extensive use of stone blades; the widespread manufacture of a variety of objects from bone, antler, ivory, and wood; the invention of new equipment, such as the spearthrower and the bow and arrow; the domestication of the dog; and the appearance of art and decoration. The Upper Paleolithic also represents an important phase in the geographic expansion of the human species. There were more sites in more places than ever before. Virtually all the earth’s diverse environments, from tropical rain forest to arctic tundra, were inhabited during this period. Africa, Europe, and Asia were filled with groups of hunter-gatherers, and Australia and North and South America were colonized for the first time.
What forces were at work making us more human? How did evolution select for more creative individuals?
y
Figure 4.26 Three views of a blade: a flake with a length at least twice its width.
_ Concept The Upper Paleolithic
123
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Figure 4.27 Upper Paleolithic manufacture of blades and blade tools. Blade manufacture is a kind of mass production of many elongated flakes. A pointed piece of bone or antler is struck with a hammerstone to remove the blade from the core using the indirect percussion technique.
_ _
Figure 4.29 Solutrean laurel leaf point produced by retouching and thinning the surface of the artifact.
124
Chapter 4 The Hunters
Page 124
Figure 4.28 Blades as forms, or blanks, for making other tools. Increasing specialization in the function of tools was an important trend during the Paleolithic. This illustration shows some of the many types of tools that were made from blades.
The archaeological materials of this period are best known from Europe, especially from southwestern France, an important hub of archaeological activities during the twentieth century. In Europe, the Upper Paleolithic replaced the Middle Paleolithic after 40,000 years ago. Excavations over the past 100 years in the deep deposits of caves and rockshelters in this area have exposed layer upon layer of materials from the last part of the Pleistocene. These excavations and studies of the contents of the layers resulted in the recognition of a sequence of Upper Paleolithic subperiods, known as the Châtelperonnian, Aurignacian, Perigordian, Solutrean, and Magdalenian (Figure 4.25). In central and eastern Europe, the Upper Paleolithic remains are designated the Gravettian, roughly equivalent to the Perigordian in the west.
The earliest skeletal remains of Homo sapiens sapiens found in western Europe date to 42,000 years ago, following the appearance of blade tools and other distinctively Upper Paleolithic artifacts. These anatomically modern individuals were originally called Cro-Magnon, after the place in France where they were first discovered. In spite of this distinctive name, they were indistinguishable from fully modern humans. Lacking the robust frame, heavy brow ridges, and protruding jaw of the Neanderthals, the H. sapiens sapiens face sits almost directly under a bulging forehead. A chin reinforces the smaller, weaker jaw and its smaller teeth. Cranial capacity is fully modern, and there is no reason to assume that Cro-Magnons were intellectually different from us. The material remains left by these Upper Paleolithic societies rein-
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
force the idea that by this time our species had indeed arrived as creative creatures. Blade-manufacturing techniques and blade tools characterize the Upper Paleolithic. Stone blades are a special form of elongated flake, with a length at least twice its width and sharp, parallel cutting edges on both sides (Figure 4.26). Blades can be mass-produced in large quantities from a single nodule of flint, removed from a core in a fashion akin to peeling a carrot (Figure 4.27). Blades also provide a form, or blank, that can be shaped (retouched) into a number of different tools. Projectile points, burins, knives, drills, and scraping tools can all be made from a basic blade form (Figure 4.28). Another distinctive aspect of Upper Paleolithic stone tool manufacture is the appearance of special flaking techniques during the Solutrean period, to make thin, beautiful, leaf-shaped points in several sizes (Figure 4.29). Some of these points were used for spears and some perhaps for arrows, while others may have served as knives. These tools are among the finest examples of the flintknapper’s skill from the entire Paleolithic. At the end of the Solutrean, however, these flaking techniques largely disappeared from the craft of stone tool manufacture, not to be used again for thousands of years. Many new kinds of tools—made of materials such as bone, wood, ivory, and antler—also distinguish the Upper Paleolithic. Spearthrowers, bows and arrows, eyed needles, harpoons, ropes, nets, oil lamps, torches, and many other things have been found. Hafting and composite tools, incorporating several different materials, were also introduced during the Upper Paleolithic. Resin and other adhesives, for example, were used to hold stone tools in bone or antler handles. Spearthrowers provide an extension of the arm, enabling hunters to fling their darts with greater force and accuracy (Figure 4.30). A hunter with a spearthrower can kill a large animal such as a deer from a distance of 15 m
Page 125
Figure 4.30
An artist’s reconstruction of a spearthrower in action.
(50 ft). Spearthrowers of bone, wood, or antler usually had three components: a handle, a balance weight, and a hook to hold the end of the spear. These spearthrowers were often elaborately decorated, with the carved figures of animals used for the weight (Figure 4.31). By the end of the Upper Paleolithic, the spearthrower was replaced by the bow and arrow as the primary hunting weapon. The bow provided an even more accurate means of delivering a long-distance, lethal blow to an animal. Dogs were domesticated during the Upper Paleolithic, probably for the purpose of hunting. As temperatures warmed at the end of the Pleistocene and the European forests spread back across the continent, woodland species of animals became more common but less visible to the hunter. A strong sense of smell, lacking in a human hunter, to locate prey was well developed in his faithful canine companion.
blade A special kind of elongated
flake with two parallel sides and a length at least twice the width of the piece. blank A basic form or preform from which various kinds of tools can be shaped.
_ Concept The Upper Paleolithic
125
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 126
Figure 4.31 One end of a decorated spearthrower from the Upper Paleolithic in France. A carving of two embracing elks provides the balance weight.
Fine bone needles with small eyes document the manufacture of sewn clothing and other equipment from animal skins. Several categories of carved artifacts—buttons, gaming pieces, pendants, necklaces, and the like—marked a new concern with personal appearance, an expression of self, and the aesthetic embellishment of everyday objects. This development was closely related to the appearance of decorative art. Figurines, cave paintings, engravings, and myriad decorations of other
_ _
126
Chapter 4 The Hunters
objects reflect the creative explosion that characterized Upper Paleolithic achievement. There is also compelling evidence for a celebration of the seasons and an awareness of time in the archaeological remains from the Upper Paleolithic. Finally, the suggestion of counting systems and the beginning of a calendar of sorts—or at least a recording of the phases of the moon— may have appeared at this time (see “Symbols and Notation,” p. 140).
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Page 127
Site
:
Dolni Vestonice Mammoth hunters in eastern Europe
The woolly mammoth of Pleistocene Europe was a magnificent creature. As seen in cave paintings and frozen remnants from Siberia, this animal had a huge domed head atop a massive body covered with long fur. The mammoth was roughly one-and-a-half times the size of a modern African elephant and must have been formidable prey for the late Pleistocene hunters of Europe. In addition to mammoths, herds of wild reindeer, horses, woolly rhinoceroses, and other species roamed the tundra of Europe. The mammoth, however, was the primary game in the east and provided the bulk of the diet for the inhabitants of this area. At one site in the Czech Republic, the remains of 800–900 mammoths have been uncovered. The remains of the camps of these mammoth hunters were fortuitously buried under deep deposits of fine silt. This silt was originally picked up by the wind at the edges of the ice sheets, carried in the air across central Europe, and gradually deposited as blankets of sediment, known as loess (pronounced “luss”). Numerous prehistoric sites were slowly covered by this airborne dust; bone, ivory, and other materials have been well preserved in it. The major problem with such sites is simply finding them, because they are hidden under very deep deposits of loess. Near the town of Dolni Vestonice (dol-NEE ves-toe-NEET-za), in the south-central part of the Czech Republic, the enormous bones of extinct mammoths were first uncovered in the course of quarrying loess soils for brickmaking (Figure 4.32). Although excavations were initially undertaken in 1924, the extent of the prehistoric occupation was not truly recognized until the commencement of the work that began in 1947 and continues today. Large horizontal excavations have removed the deep loess deposit covering
the site and exposed a large area containing dwelling structures, mammoth bones, and many intriguing artifacts dating to about 25,000 years ago. During the late Pleistocene, the area was one of tundra and permafrost, situated north of the treeline in Europe. Little wood was available, except possibly for small stands of willow and other species in sheltered valleys. Broad expanses of grass, moss, and lichen provided food for the herds
Figure 4.32 Excavations at the site of Dolni Vestonice, uncovering the piles of mammoth bones on the south side of the site. Notice the deep layer of loess deposits above the level of excavation.
loess Wind-blown silt deposited in deep layers in certain parts of the Northern Hemisphere.
Site Dolni Vestonice
127
_
pri31057_ch04.pp104-163.qxd
10/1/09
Figure 4.33 A residential structure from Dolni Vestonice showing the semisubterranean floor, with flat stones, fireplaces, and mammoth bones used for the framework. Postholes are indicated by small dark circles, and fireplaces are marked by black diagonal lines. The bone framework was likely covered with animal hides to complete the structure. The hut is approximately 6 m in diameter.
solifluction A phenomenon in which
freezing and thawing of the ground results in slippage of the surface. red ochre An iron mineral that occurs
_ _
in nature, used by prehistoric peoples in powdered form as a pigment for tanning animal skins.
128
Chapter 4 The Hunters
7:54 AM
Page 128
of mammoths, horses, and reindeer that were the predominant fauna of the area. The permafrost was responsible for large-scale movement of the ground surface, a phenomenon known as solifluction. Alternate freezing and thawing of the ground resulted in the disturbance of many of the remains at Dolni Vestonice. For this reason, it is somewhat difficult to interpret evidence from the site. The highest layers in the deposits, containing a campsite, are still reasonably well preserved. This camp lay on a projecting tongue of land, along a local stream that becomes a bog just at the eastern edge of the site. Part of the site sits on a ridge, providing a good view of the valley of the nearby Dyje River. The effectiveness of the mammoth hunters is dramatically portrayed in the scatters of mammoth bones marking the boundaries of the settlement. The bones of at least 100 mammoths were piled up in an area measuring 12 × 45 m (40 × 140 ft). Stone tools and broken bones suggest that this was a zone where animals, or parts of animals, were butchered and where skins may have been cleaned and prepared. Other piles of bones were found throughout the settlement, often sorted according to kind of bone, presumably for use as fuel and raw material for construction. Fires were lit on some of these bone
piles, as evidenced by ash, perhaps as a defense against predatory animals. Stones, earth, wooden posts, and mammoth bones were used in the construction of structures at the site (Figure 4.33). The first structure to be uncovered was a very large oval, 9 × 15 m (30 × 50 ft), with five regularly spaced fireplaces inside. The size of the structure, about half a tennis court, and its contents suggested to the excavators that this was an open windbreak, without a ceiling, rather than a roofed structure, and that it was occupied primarily during summer. The wall posts were supported with limestone blocks and were likely covered with animal hides. At least three roofed huts have been found in this area. These structures are partially dug into the loess; they contain one or two hearths and have numerous large mammoth bones on top of the floor. These bones are probably the remains of the framework for the roof, which would have collapsed onto the floor of the structure after the site was abandoned. In an open area near the center of the compound was a large hearth, almost 1 m (3 ft) deep and several meters (about 10 ft) across, which may have been a common, central fire for the community. In the ashes of this fire, an ivory carving of a female figure, called the Venus of Vestonice, was found (see “Portable Art,” p. 136). Another structure, uncovered in 1951, was found some 80 m (250 ft) along the stream to the west of the main concentration (Figure 4.34). This structure was smaller, 6 m (20 ft) in diameter, and very unusual. The floor of the hut had been dug into the loess slope to level it and to provide more protection against the elements. Limestone blocks were placed against the excavated slope to buttress the wall. Posts were also supported by these blocks at the front of the hut. Hollow bird bones were found inside; they were cut at the ends and may have functioned as musical instruments. In the center of the hut was an ovenlike fireplace with a domed clay structure raised around it. The oven was made of fire-hardened earth and ground
pri31057_ch04.pp104-163.qxd
10/1/09
7:54 AM
Wood poles
Page 129
Animal hides
Interlaced reindeer antlers
Large animal bones and boulders holding hides in place
Large animal bones for structural support
Figure 4.35 An ivory carving of a mammoth from Vogelherd, Germany (8.8 cm long).
Semisubterranean floor Cross section/interior
Entrance Exterior
Figure 4.34 A reconstruction of a mammoth-bone structure from the 18,000-year-old site of Mal’ta in south-central Russia, which may closely resemble the structures at Dolni Vestonice. Bones and animal skins were used as construction materials.
limestone. In the deep pile of ashes and waste that stood on the floor of the hut were found more than 2300 small clay figurines that must have been fired in the oven. This is the earliest example of the use of fired clay in the world, some 15,000 years before the invention of pottery. The figurines consisted of heads, feet, and other fragments of animal effigies and fired lumps of clay. Even the fingerprints of the maker were preserved in some of the pieces. The depth and extent of deposits at Dolni Vestonice, along with the presence of both summer and winter huts, suggest that this site may have been occupied throughout the year. The remarkable artifacts and other materials found at the site confirm the impression that this is an unusual site indeed. Flint tools at the site belong to the Gravettian, the Upper Paleolithic of eastern Europe. Tools are made from narrow blades in the form of points, knives, burins, and others. There are also numerous tools made from mammoth bone and ivory: awls, needles, knives, spearpoints, lances, and digging implements. Ornaments in the form of pendants, necklaces, headbands, and the like are made of carved bone, ivory, and shell. Some of the shells were from the Mediterranean Sea, several hundred
kilometers to the south, indicating either travel or trade. Other objects carved of antler or ivory, or made of baked clay, have no clear practical purpose and probably served as ritual objects in the ceremonies that took place at the site (Figure 4.35). Perhaps the most remarkable finds may involve two representations of the same individual. Excavations in 1936 uncovered a small ivory plaque about 4.6 cm (1.8 in) high, with a crudely incised human face portrayed on it. The face is asymmetrical, with the left eye and the left half of the lip somewhat lower than the right. A second carved ivory head was found in 1948 in the open summer hut. This three-dimensional head also portrays an individual, and the left side of the face is somewhat distorted and asymmetrical (Figure 4.36). Finally, a burial was excavated in 1949, discovered beneath two huge shoulder blades from a mammoth. The skeleton belonged to a woman and was covered with red ochre, and a flint point was buried near her head (Figure 4.37). A study of the facial bones of this individual showed that she suffered from partial paralysis of the left side of her face. It seems entirely possible that the two faces carved in ivory are representations of the person in the grave.
Figure 4.36 The head of a woman with an asymmetrical face, carved in ivory (4.6 cm high).
Figure 4.37 The buried skeleton of an elderly female from Dolni Vestonice. The bones of the left side of the face revealed congenital nerve damage, probably resulting in an asymmetrical facial expression. The woman was buried under two mammoth shoulder blades, and the bones of an arctic fox were found next to her in the grave. Site Dolni Vestonice
129
_
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 130
Site
:
The Cave of Lascaux A monument to human creativity
Figure 4.38 The Upper Paleolithic cave of Lascaux in southwestern France contains many paintings of a variety of animals. This running horse is depicted in two colors along with two feathered darts or plants.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on some of the best examples of cave art in the world, see the Internet exercises on your online learning center.
_ _
130
Chapter 4 The Hunters
Nestled in the lovely countryside of southwestern France, the Vezere (VEZair) River runs past some of the most important Paleolithic sites in the world. This area, known as the Perigord (pear-e-GORE), is a prehistorian’s dream. The spectacular landscape contains not only hundreds of important archaeological sites but also the kitchens and cellars of many superb French chefs. The limestone plateau is dissected by numerous streams and rivers that have carved high cliffs along the courses of the valleys. These cliffs contain caves and rockshelters, which provided residence for generations of Paleolithic groups. Over time, the entrances to many of these caves have collapsed and hidden the caverns completely. The cave
of Lascaux (lahss-CO) was discovered by chance in 1940 when a young boy noticed a hole in the ground where a pine tree had been uprooted by the wind. After dropping some rocks into the hole, he and several friends slid down the opening and into a cave, which we now know contains the most important collection of Upper Paleolithic art in the world (Figure 4.38). Lascaux had been sealed for perhaps 15,000 years, the outside world completely unaware of the splendor it held. After World War II, the cave was opened as an underground museum. For two decades, it was one of the major tourist attractions in France, receiving as many as 1000 visitors a day. Unfortunately, the flow of tourists changed the environment of the cave
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
G L A C I A L
0
I C E
300 km
ATLANTIC OCEAN
FRANCE CAN
TA
BR
IAN
Lascaux Chauvet M
TS .
PY
RE
N EE
Cosquer
S
dramatically, raising the temperature and humidity and bringing in dust. These changes caused the growth of a fungus on the walls that began to cover and eventually flake off parts of the paintings. The cave was closed in 1963, and efforts were begun to halt the spread of the fungus, remove it, and preserve the paintings for posterity. The problem was fixed, but the cave has not been reopened to the public. Instead, the French government built a duplicate of the cave, Lascaux II, where tourists can view copies of the beautiful animal paintings that grace the walls of the original galleries. The artwork of the Upper Paleolithic can be divided into two major categories: (1) mural art—paintings and engravings on the walls of caves—and (2) portable art—carvings, figurines, and other shaped or decorated pieces that can be moved from place to place. Upper Paleolithic mural art is found primarily in France and Spain (Figure 4.39), although ancient examples have also been found in South Africa and Australia. Portable art is found throughout Europe and much of the rest of the Old World. Although Upper Paleolithic mural art was also painted at cave entrances and along cliff faces and rock outcroppings, only deep inside the caves of France and Spain has it survived the erosive forces of nature.
Page 131
SPAIN
Mediterranean Sea
Figure 4.39 Locations of the western European painted cave sites. The painted caves are primarily in the Perigord region of southwestern France, in the Pyrenees mountains between France and Spain, and in the Cantabrian Mountains of northern Spain, as indicated by dots on the map. The map also shows the extent of continental glaciation in northern Europe during the coldest period of the Upper Paleolithic.
Figure 4.40 Some of the human heads engraved on the walls of the cave at La Vache.
mural art Painting, engraving, and
sculpting on the walls of caves, shelters, and cliffs. portable art Decorated materials that can be moved or carried.
Site The Cave of Lascaux
131
_
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 132
Figure 4.41 Three depictions on the walls of the cave of Les Trois Frères, France, showing humans in animal costume. These individuals may be dancers, sorcerers, or hunters.
y
_ _
Many of the new aspects of the Upper Paleolithic are related to depictions of nature, especially the animal world. Why would it have been important to make images of these creatures?
Figure 4.42 The number and kinds of hand depictions appearing on the walls of the Upper Paleolithic cave of Gargas in France. Note the many missing fingers and digits.
132
Chapter 4 The Hunters
The cave interiors were not living areas; they were visited only briefly by the artists and other members of society. The paintings are almost exclusively of animals; humans are only rarely represented. Human figures are depicted more commonly in engravings, both on cave walls and in portable objects (Figure 4.40). Some of the engravings of humans seem to show men wearing the skins and heads of animals (Figure 4.41). These individuals may have been dancers, participants in some ceremony, or camouflaged hunters. Single human hands are outlined on many of the cave walls. Curiously, they are often depicted without knuckles or even entire fingers. For example, at the cave of Gargas in the French Pyrenees, all but 10 of 217 hand paintings have missing fingers (Figure 4.42). These hands may be simple signatures or graffiti but their exact meaning is unknown. The cave paintings themselves are rendered in outline and often colored in monochrome or polychrome. Animals are most often depicted in profile. The paintings are sometimes located high on the ceilings in the darkest areas of the caves; light and some form
of scaffolding would have been needed to paint in these places. Pieces of rope and pine torches have been found in some caves, along with simple oil lamps made of stone bowls. Over 150 fragments of such lamps have been recovered at Lascaux. The paintings must have had a great impact viewed in the light of a flickering torch by awed individuals deep in the earth’s interior. The cave art is, by and large, carefully planned and skillfully executed, capturing both the movement and the power of the animals that are rendered. It is not graffiti, nor is it hastily sketched. The quality of the paintings is often such that we must assume there were recognized artists in the societies of the Upper Paleolithic. Most of the paintings at Lascaux date to around 17,000 years ago, during the Magdalenian period. The art is dated from fragments of paints and other artifacts found in archaeological layers of known age in the cave. The paints were a blend of mineral pigments mixed with cave water, and the chewed end of a stick or pieces of hair or fur were used to tamp the paint onto the walls. The common colors in the paintings are black (charcoal and manganese oxide), yellow and red (iron oxides and clay), and occasionally white (clay and calcite). Lascaux is a lengthy, narrow chamber a little longer than a football field. In spite of its relatively small size, more than 600 paintings and 1500 engravings grace the walls of the cave, making it the most decorated of the magnificent painted caves in Spain and France (Figure 4.43). The opening of the cave leads into a large chamber, some 20 m (65 ft) long, filled with huge animal paintings. Four large bulls, up to 5 m (16 ft) long, stride across the ceiling. In the adjacent halls and passageways, hundreds of paintings depict bison, deer, horses, wild cattle, and other animals. Very specific characteristics, such as spring molting, are shown in some of the paintings. Animals are often depicted as pregnant or with their meaty haunches exaggerated. In several instances, feathered
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 133
6 Apse
Passage Rotunda of the Bulls
9
8
7 2
Cavern of the Cats 12
11
10 0
10 m
1
3 5 4
darts are heading toward the animals. Certain abstract patterns also appear either in isolation or in association with animals. Rows of dots and multicolor checkerboard patterns are painted at various places in the cave. Most of the paintings show one animal or a group of animals; there is little attempt at scenery or storytelling. Many of the paintings are superimposed over older ones, with little apparent regard for the previous work of other artists. Large herbivores, which provided much of the meat for Upper Paleolithic hunter-gatherers, most frequently appear in large chambers and open areas in the caves. Curiously, however, the most important game animal at this time, the reindeer, appears only once at Lascaux. Such dangerous animals as carnivores, bears, and rhinoceroses more often are found in the deep recesses of the cave and far-removed crevices. One of the most remarkable paintings is found in a narrow, 5-m (16-ft) shaft off to the side in the cave (Figure 4.44). At the bottom of this shaft, a large woolly rhino with raised tail faces to the left, a series of dots near its
hindquarters. Across a small crevice appears a striking scene of beast and man, the only human figure painted in the cave. A beautiful multicolor bison on the right is mortally wounded, its entrails spilled by a spear. The dying animal is either down on the ground or charging the human figure on the left. This figure, obviously male, is shown in mere outline, depicted with a birdlike face. On the ground nearby lies a long object with several barbs, perhaps a spearthrower, and a bird with a single long leg, possibly an important symbol. Is this painting a memorial to a hunter, a member of the bird clan or totem, killed by the bison? Does the rhino play a role in the scene? Such questions point out the difficulties involved in trying to read the minds of prehistoric people. We can speculate about, but we cannot know, what was intended by these paintings. There are several schools of thought on the meaning of the cave paintings from the Upper Paleolithic. An apparent emphasis on pregnant animals has often been interpreted to represent a concern with fertility and the bounty of nature, reflecting an awareness of the
Figure 4.43 The distribution of animals and designs at Lascaux. The numbers indicate major areas of art in the cave. Large prey animals tend to be in the major galleries of the caves; carnivores and other depictions are in less-accessible areas.
_ Site The Cave of Lascaux
133
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 134
Figure 4.44 Man and bison at Lascaux. This painting represents one of the very few examples of storytelling in the art of Lascaux, and it is the only painting of a human in the cave. The depiction is at the bottom of a deep, narrow shaft and has four elements: a detailed color drawing of a bison mortally wounded by a spear, a black outline of a male human with a birdlike face, a bird on a stick beneath the human figure, and a spearthrower with hook and handle lying on the ground beneath the man. The painting is subject to various interpretations, ranging from a memorial to the death of a kinsman to the depiction of an Upper Paleolithic myth.
_ _
No matter how many caves one has explored, no matter how magnificent or crude or abstract the figures, it always comes with a catch of breath. It may be a bull or a bison drawn larger than life or an engraved horse no bigger than your little finger. It may appear high in a fissure or down close to the floor, on wide exposed surfaces for all to see or in private crawl-in places, painted bare red or black outline or in rich polychrome, starkly grand or delicate in a low key—a variety of locations and styles, and yet all part of a single tradition that endured for some twenty millenniums. —John Pfeiffer (1982)
134
Chapter 4 The Hunters
importance of reproduction and the replenishment of the herds on which these people depended for food. Other scholars, pointing to the exaggerated hips and haunches of the animals and spears in flight, argue for a concern with the hunting of animals for meat. Hunting rites and ritual killings of animals before a hunt might magically help ensure success in the quest for food. A few prehistorians suggest that the cave paintings were simply “art for art’s sake,” a means for artists to express themselves and to change the way their fellow humans saw the world. Still others suggest that the painted caves were primitive temples, sanctuaries for ceremony and ritual, such as the initiation of the young into society. Huge animals flickering in the light of torches and lamps deep within the bowels of the earth would have provided a breathtaking experience for the uninitiated. Footprints preserved in the muddy floors of painted caves in France indicate that people of all sizes walked in
the cave. Margaret Conkey, of the University of California at Berkeley, argues that these caves may have served as a focus of social activity for large groups of people. She suggests that the caves may have been a permanent symbol on the landscape and a place for the ceremonies and rituals associated with the aggregation of several different groups of hunter-gatherers. More than 200 painted caves have been discovered in France over the past 100 years. In the past decades, several major new art sites have been revealed (Figure 4.45). The cave of Chauvet (SHOW-vay) was found in a tributary of the Rhône River in the south of France in 1995, containing more than 300 paintings and engravings. The site was discovered by French spelunkers who cleared a small hole at the surface and climbed down 9 m (30 ft) into a great chamber. The cave is at least five times larger than Lascaux. There are several groupings of animals, including bears and rhinos, on the walls, in
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 135
addition to a number of solitary animals. New radiocarbon dating of some of the art in the cave places it at 36,000 years old, making these some of the oldest known paintings in the world (see “Radiocarbon Dating,” p. 145). Another important cave, Cosquer (KOS-care), was found underwater by divers off the Mediterranean coast of France, near Marseilles, in 1992. Swimming through the long, narrow opening of this cave, divers entered dry chambers containing the untouched remains of Upper Paleolithic people, footprints, lamps, torches, hearths, and the like, along with many extraordinary paintings. Rising sea levels at the end of the last glaciation submerged the entrance of the cave. The opening of this cave, now almost 40 m (120 ft) below sea level, was along the shore during the Upper Paleolithic, when the sea level was lower. A number of the paintings reflect the proximity of the Mediterranean shore, depicting seals and seabirds such as the great auk. Most of the art in the cave dates to around 18,500 B.P. The magnificent art of the Upper Paleolithic represents an initial awakening of the creative spirit, an explosion of our aesthetic senses. Such a transformation may also signify major changes in the minds of Upper Paleolithic people and/or in the way they viewed the world and organized their lives and their society.
© The New Yorker Collection 1980 Lee Lorenz from cartoonbank.com. All Rights Reserved.
Figure 4.45 Cave paintings cover the walls at the site of Vallon-Pontd’Arc, discovered recently in southern France.
“Does it strike anyone as weird that none of the great painters have been men?”
_ Site The Cave of Lascaux
135
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 136
Concept
<
Portable Art A sense of design and beauty The aesthetic sense that appeared during the Upper Paleolithic with the arrival of fully modern humans was expressed in a variety of forms. Carving, sculpting, and molding of various materials, including clay, antler, wood, ivory, and stone, is evidenced throughout this period (Figure 4.46). The decoration of artifacts and other objects occurred throughout the Upper
Figure 4.46 Examples of Upper Paleolithic decoration of bone, antler, ivory, and wood objects. Such decoration was applied to a variety of pieces, some utilitarian and others more symbolic.
_ _
136
Chapter 4 The Hunters
Paleolithic beginning about 35,000 years ago. There is remarkably little evidence for the nonpractical modification of equipment and utensils before the appearance of Homo sapiens sapiens in Europe. Only a handful of decorated objects have been found in Middle Paleolithic contexts. Beginning in the Aurignacian some 35,000 years ago, however, bone be-
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
came a common material for human use, modification, and decoration. For example, a variety of bone points date to this early period. Initially simple and plain, such points had become heavily barbed and decorated by the end of the Upper Paleolithic. At the same time, carved bone and antler figurines of both humans and animals began to appear in the archaeological record. Perhaps the most spectacular portable objects from the Upper Paleolithic are the “Venus figurines.” These small sculptures appeared throughout most of Europe during a brief time around 25,000 years ago. The figures were engraved in relief on the walls of caves; carved in the round from ivory, wood, and stone such as steatite; and modeled in clay. The female characteristics of these statuettes are usually exaggerated: breasts, hips, buttocks, and thighs are very large; the head, arms, hands, legs, and feet are shown only schematically (Figure 4.47). The pubic triangle is sometimes outlined; one figurine has a detailed vulva. Some of the figurines appear to be pregnant, and others are displayed holding a horn, perhaps a cornucopia or horn of plenty, to imply fertility, bounty, and reproduction. Probably 80% of the prehistoric art known today comes from the last stage of the Upper Paleolithic, the Magdalenian. Objects with a short life were decorated in a cursory fashion, whereas more important pieces with a longer life expectancy were heavily ornamented. Spearthrowers were decorated elaborately, with carved animals serving as the counterweight and end-hook. Engraved bone was common, and such portable art was often painted as well. Body adornments, including necklaces, bracelets, and pendants, also appeared in the Magdalenian. Portable art was more common in the larger settlements than in smaller ones. This pattern suggests a connection between art and the ritual activities that likely occurred when larger groups of people came together. Huntergatherers commonly aggregated in a larger group at a certain time each year
Page 137
to exchange raw materials and learn new information, to find mates, and to celebrate important events, such as marriage and initiation into adulthood. Rituals and ceremonies provided a common bond in both the physical and psychic realms; dance, trance, and the reaffirmation of common beliefs were important aspects of such gatherings. Decorations in the form of masks, face and body painting, costumes, and the like were probably used during such ceremonial occasions (see “Contemporary HunterGatherers,” Chapter 5, p. 194).
Figure 4.47 Venus figurines in various shapes and sizes. The individual holding the horn or cornucopia (top row, far right) supports the interpretation of these figures as symbols of fertility, nature’s bounty, or “Mother Earth.”
One of the more intriguing features of becoming fully modern is the evidence for jewelry and self-adornment. Artifacts such as beads, pendants, and decorated batons suggest that individuals were distinct, that egos were emerging. Why is such behavior an important part of being human?
Concept Portable Art
y
137
_
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 138
Site
:
Pincevent Brief stops by reindeer hunters in the Upper Paleolithic Several shallow fords on the Seine River in northern France were important crossing points for migrating reindeer herds when tundra and permafrost covered northwestern Europe at the end of the Pleistocene. The sandy banks and bars of the Seine and other rivers were the sites of the camps of Magdalenian reindeer hunters approximately 12,000 years ago. In the vicinity of Paris, the Seine may lie near the northern end of what was a major route of reindeer migration at the close of the Pleistocene, with herds moving north each spring from the south of France. Scatters of stone, bone, antler, hearthstones, and charcoal mark these ephemeral summer encampments of reindeer hunters. A number of such sites, including one called Pincevent (PONCE-von), were quickly buried and extremely well preserved (Figure 4.48). The archaeological remains are found in thin layers of clay, deposited when the river flooded annually. The river floods must have been gentle, because there is little disturbance of the materials. Several artifacts were found standing upright, and two crushed bird eggs remained at the site. At Pincevent, at least four levels with archaeological remains have been
Figure 4.48 Two hearths at Pincevent, France, with the distribution of stones and bones deposited here some 12,000 years ago. All these objects lie where they were discarded by the Upper Paleolithic hunters.
_ _
138
Chapter 4 The Hunters
recognized, extending over an area of 2 ha (5 acres), larger than a soccer field. Excavations, originally directed by A. Leroi-Gourhan and M. Brezillon, began in 1964 and continue today. The excavators intentionally exposed broad horizontal areas of the site, leaving features, artifacts, and bones in place. In this way, entire “living floors” could be seen and the pattern of discarded materials studied to determine where people slept, cooked, made tools, and so on. The excavators also made latex rubber casts of many areas, which were then painted to reconstruct and permanently preserve the archaeological remains. The concentrations of materials average 60–70 sq m (650–750 sq ft), about one-half of one side of a tennis court, and they probably represent single tents or structures as the residence and focus of activity of a few hunters. Each concentration contains 20,000–30,000 stone artifacts and other materials. Stone blades, a major product at these sites, had a variety of purposes. At the nearby site of Etiolles, extremely long blades were produced, some more than 80 cm (30 in) long, from large nodules of raw material weighing 40 kg (90 lb) or more. The
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
blades from these nodules were as long as 50 cm (20 in). The absence of wear marks on the edges of some of these blades suggests they had not been used and were being stockpiled for some later purpose. At Pincevent, one of the most important areas excavated to date is in Layer IV. This area contains 94,000 kg (200,000 lb) of flint artifacts, the skeletal remains of at least 43 reindeer, firecracked rock, ochre, and several shallow pits and fireplaces (Figure 4.49). Red ochre stains are concentrated around three large fireplaces. The excavators suggest that activities were centered on three contemporary huts, each with an associated fireplace (Figure 4.50). Each hut contained a central zone for actual living space and surrounding zones of domestic activities and refuse disposal. The intensity of activity decreased with distance from the hearths. Small piles of waste materials from stone tool manufacture lie on one side of the hearths, finished tools and red ochre on the other. Near one of the hearths is a large stone that was likely the seat of a flintworker. Most of the flint was available in the immediate area of the site. A few pieces, however, came from some distance, confirming the mobility of the hunters who stopped here. Reconstruction or refitting of the pieces removed from the flint nodules provides a good indication of how tools were made. Moreover, pieces that are missing and not found at the site provide evidence of which tools were carried elsewhere. Finally, the scattered locations of pieces that fit together indicate how the tools and waste materials were moved about at the site. Most of the reindeer at the site were killed and butchered during summer. The distribution of bones on the living floor is similar to that of the flint debris. Larger bones were at the periphery; smaller pieces and smaller fragments were found near the fireplaces. The bones from a meal were apparently tossed away from the hearth. Small fragments of antler were found near the hearths, but larger pieces were discarded at the edge of
Page 139
Figure 4.49 The distribution of stones and bones around two hearths at Pincevent. The large stones were used for sitting during stone tool manufacture; the scatters of lithic debris were produced as a result.
a b
the activity zone. Most antler working was apparently done at this periphery. The lack of sweeping or cleaning of the living area suggests that the occupation at Pincevent was very brief. Lewis Binford, of Southern Methodist University, has questioned the existence of actual tents or structures at Pincevent, arguing that the distribution of materials observed in arcs of debris around the hearths could just as easily have happened without tents or huts present. There is no definitive evidence that the hunters built shelters on the site. Binford also argues that all the materials around the two largest hearths could have resulted from one or two individuals working and shifting position in response to wind direction. Binford based his suggestions on observations of Eskimo hunters in Alaska and the manner in which they moved away from the smoke of a fire on a windy day.
Figure 4.50 Reconstruction of a possible hut around one of the Pincevent hearths: (a) concentration of artifacts, (b) hearth.
_ Site Pincevent
139
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 140
Concept
<
Symbols and Notation Evidence of seasonal awareness, numbers, and phases of the moon Some of the many decorated objects from the Upper Paleolithic contain unusual images that are not easy to understand. These designs are carved into the polished surface of bone using pointed stone tools. The motifs that often occur together suggest that specific concepts were being depicted. One example is a bone knife from the French site of La Vache (la VASH). The design on one side of this piece has two animal heads—a doe and an ibex—wavy lines that may represent water, and three plants. The other side of the knife shows the head of a bison in autumn rut, four plant motifs that
Figure 4.51 An engraved bone piece from the site of La Vache, France. Two horses appear at the bottom of the piece, one with a number of arrows drawn into the animal. The repeated sets of tally marks, usually in groups of 8–10, were made with different stone tools, suggesting that something was being counted or recorded over time.
_ _
140
Chapter 4 The Hunters
may be pine branches, a drooping stem, and three seeds or nuts. Alexander Marshack (1918–2004) suggested that the two sides of the knife are intended to convey images of spring and fall, in recognition of the seasons and their distinctive characteristics. Other bone artifacts have unusual combinations of notches and patterns of dots that are more difficult to comprehend. One example of such an object comes from the cave of La Vache and dates to about 14,000 years ago. The polished fragment of long bone, about 15 cm (6 in) long, is decorated at the base with one entire horse and the head of another (Figure 4.51). About 10 pointed lines are drawn into the complete horse, perhaps symbolically representing the hunt. When the bone with the horses broke, it was reused to flake flint tools; later, elaborate sets of marks were added in a series of rows on both sides of the bone above the horses. Marshack had undertaken detailed, microscopic studies of these objects to determine how the notches and dots were placed on the bone. He was able to distinguish both the type of pointed tool that made the marks and the order in which at least some of the marks were made. For example, at least four tools were used to carve the arrows in the horse on the La Vache bone; the tally marks all occur in sets or blocks, each of which was made with a different tool in a slightly different shape. The number and pattern of the marks suggest that they were added over a period of time. On other decorated bones from the Upper Paleolithic, Marshack observed marks in groups of 30 or 31. The number 7 also seems to regularly define groups of marks on other objects. Although such marks on bones have been considered decoration, they could signify some kind of tally, counting a series
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 141
Figure 4.52 The 32,000-year-old bone plaque from Abri Blanchard, France (top), with the area of engravings shown in detail (bottom). Marshack suggests that these marks record the phases of the moon. The different shapes and colors on the drawing of the piece indicate marks made by various stone tools, perhaps indicating the passage of time.
0
of events, or perhaps the number of hunting kills. Possibly the most intriguing decorated bone object yet found is from the site of Abri Blanchard in France. A flat, irregular rectangle, this bone has no animal or figure engravings, but rather a series of carved and engraved notches and marks (Figure 4.52). The 80 notches cover about half the edge of the object, and the marks form a semicircular pattern of two parallel lines on the flat surface. Microscopic examination indicates that at least 24 tools were used to make the 69 marks on the surface of this bone. Marshack believed that these marks record the phases of the moon. The shape of the marks changes with the moon’s phases, over a period of about 6 months. Marshack argued that our Upper Paleolithic forebears were noting the passage of time, reckoning the year
2 cm
according to the seasons and a lunar calendar. Marshack’s discoveries are controversial, but the idea that Upper Paleolithic people were capable of counting and notation, as well as symbolic representation, does not seem farfetched, in light of the other evidence of their creativity and accomplishments.
_ Concept Symbols and Notation
141
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 142
Site
:
Lake Mungo, Australia The spread of Homo sapiens sapiens Although it may seem strange to find Australia in the middle of this discussion of the Upper Paleolithic, there are several important things to be learned from a consideration of the evidence from the Land Down Under. Australia was colonized by Homo sapiens sapiens around 40,000 years ago. Rafts or boats of some kind were likely used. These first emigrants somehow crossed a body of water at least 100 km (65 mi) wide, far beyond the sight of land, to reach the island continent. There are archaeological sites at least 20,000 years old in all corners of the continent; the oldest, on the Upper Swan River, dates to around 38,000 years ago. At this time, during one of the coldest periods of the Pleistocene, sea level was as much as 150 m (400 ft) lower than it is today. The continental shelves were exposed, and land bridges
_ _
142
Chapter 4 The Hunters
0
SOUTHEAST ASIA
1000 km
NEW GUINEA
ine
e
L
Sahul Str ait
PACIFIC OCEAN
ac
Figure 4.53 Australia and Southeast Asia during the colder periods of the Pleistocene. At that time, sea level was as much as 150 m lower than it is today, and much of the continental shelf was exposed as dry land. The outlines of the continents changed considerably, and many of today’s islands became part of the mainland. Australia and New Guinea joined together but were never part of Southeast Asia. One of the deepest bodies of water in the world, the Sahul Strait, would have always been sea. Thus, the early inhabitants of Australia would have crossed a large body of water to reach the continent. The Wallace Line (red line) marks the divide between Asian and Australian animal species.
connected several areas formerly separated by the sea. Australia was connected to New Guinea and Tasmania, constituting a larger continent called Sahul (Figure 4.53). The Sahul Strait, a body of water that today is several thousand meters (more than a mile) deep, lay between Australia and Asia. At that time, Southeast Asia was a bridge of land connecting the mainland and Indonesia and Borneo. The first inhabitants crossed this deep, wide body of water to reach Australia. The absence of a land bridge between Australia and mainland Asia is evidenced by the fact that the animal species in the two areas are so very different. Asian placental mammals and Australian marsupial mammals have not been in contact in the past several million years. This difference was originally noted by British naturalist
W
INDIAN OCEAN
a ll
AUSTRALIA
Lake Mungo
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Alfred Russel Wallace, a contemporary of Darwin’s, and the term Wallace Line is used to designate the divide between the two distinct groups of animal species. Some of the oldest sites in Australia lie along the margins of dry lakebeds in the southeastern part of the country, an area known as Lake Mungo. Sites at Lake Mungo contain hearths and shells dating to around 32,000 years ago. This area would have contained a series of lakes with fertile shorelines in the period of initial occupation, when rainfall was higher than it is today. The sites were discovered in 1968 by a geologist who found human bones buried in a sand dune that was at least 20,000 years old. The bones appeared to have been buried at a time when the dune was active on the shore of a former lake. Further examination of the area around the bones on the dune revealed a series of stone artifacts and several patches of charcoal, which must have been the locations of hearths. Most of the hearths contained fish and mammal bones. Bird bones, eggshells, and shells from freshwater mollusks were also found in a few fireplaces. Other sites around the fossil lakeshore have revealed concentrations of shellfish, burned areas with charcoal and fired clay lumps, probably used as cooking stones. The material culture of these early inhabitants included both bone tools and stone tools, with a large number of heavy core and pebble artifacts. Several burials have been found, one of which, the remains of a woman 20–25 years old, is the oldest example of cremation yet known in the world. Other remains include another female and a male. Red ochre was used in some of the graves to cover a portion of the remains. All these individuals are Homo sapiens sapiens and document the presence of fully modern humans in the eastern part of Eurasia, several thousand years before their appearance in Europe. These early inhabitants rapidly occupied all of Australia, as indicated by the spread of radiocarbon dates across the continent. Some of the oldest rock
Page 143
art and wall paintings in the world are known from Australia (Figure 4.54). What is almost equally remarkable, however, is how little change took place here over thousands of years. Foraging was apparently a very successful and stable activity in prehistoric Australia. Hunter-gatherers arrived here almost 40,000 years ago, and they were still present when Captain Cook “discovered” the continent 200 years ago.
Figure 4.54 Australian Aborigine art begins to appear about the same time as Upper Paleolithic art in Europe. This photo shows petroglyphs (carvings on stone).
_ Site Lake Mungo, Australia
143
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 144
Concept
<
The Peopling of the Pacific Overcoming formidable obstacles
Figure 4.55 Decorated Lapita ceramic sherds.
Figure 4.56 Map of Pacific sites showing the distribution of the Lapita cultural complex. Dates are for first settlement.
Australia, Tasmania, New Guinea, and a few islands north of New Guinea were colonized some 30,000–40,000 years ago, but the open seas of much of the rest of the Pacific remained a formidable obstacle to human expansion until 3000 years or so ago. The Pacific Ocean covers one-third of the surface of the planet, some 181 million sq km (70 million sq mi), and contains half of the liquid water on earth. It is an enormous region, dotted with a few groups of tiny islands that have always been separated by thousands of kilometers. How and why were these islands discovered and settled? The colonization of the Pacific islands is marked by the appearance of distinctively decorated Lapita pottery, brought and made by the first settlers (Figure 4.55). The pottery is named after a site where it was found in New Caledonia, but its origins are ultimately in Southeast Asia. The Lapita peoples were skilled seafarers who, beginning around 1500 B.C., spread eastward
through the islands of Melanesia and into the remote archipelagos of the central and eastern Pacific (Figure 4.56). These sailors and farmers reached Tonga and Samoa by roughly 1000 B.C., the Marquesas by A.D. 300, and Hawaii by A.D. 600. Easter Island, one of the more remote places on earth, lies 1000 km from any other island and more than 2000 km from the coast of Chile. Easter Island was colonized around A.D. 400 by the people who eventually raised hundreds of colossal stone heads on the island. New Zealand seems to have been bypassed in the expansion and was not colonized until A.D. 1000. The crossing of this great ocean and the settling of the many islands by sailors and farmers is a remarkable story, one that repeatedly raises the questions of how and why. Many explanations have been offered, including new crops, new kinds of canoes, new navigation skills, wanderlust, and others, but the questions remain unresolved.
J A PA N CHINA
Hawaii (A.D. 600)
MICRONESIA Caroline Is.
BORNEO Sulawesi
INDO
NEW GUINEA Bismarck (over 40,000 B.P.) Archipelago New Britain (30,000 B.P.)
East Timor
NESIA
Kuk Huon Peninsula
AUSTRALIA (over 36,000 B.P.)
_ _
Solomon Is. Trobriand Is. Santa Cruz Is. MELANESIA Tongoa Fiji (c. 2000 B.C.) (Vanuatu) Lapita (c. 900 B.C.) (New Caledonia) (2550 B.C.)
TASMANIA (32,000 B.P.)
144
Chapter 4 The Hunters
NEW ZEALAND (A.D. 1000)
Marquesas (A.D. 500) Samoa (1800 B.C.) Tonga
Cook Is.
Raiatea Tahiti (A.D. 600)
Tuamotu Is. Society Is. Austral Is.
POLYNESIA
Easter Is. (A.D. 500)
0
Lapita Culture
0
2000 3000
4000 mi 6000 km
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 145
Concept
<
Radiocarbon Dating Absolute dates for the past 40,000 years
The Manhattan Project was the most secret and expensive weaponsdevelopment effort of the U.S. government during World War II. Laboratories in Chicago, New York, Tennessee, New Mexico, and elsewhere employed thousands of scientists and technicians to develop the first atomic bomb. Many unknown details about radioactivity were revealed for the first time during this work, such as the discovery that some radioactive elements could be used to measure the passage of time. Physicist Willard Libby, of the University of Chicago, announced the first age determinations from radioactive carbon in 1949 and received the Nobel Prize for his discovery. The key to this procedure, known as radiocarbon dating, lies in the halflife of radioactive elements. As explained in Chapter 2, many elements have both stable and radioactive atomic forms, called isotopes. Some radioactive isotopes, such as potassium, have an extremely long half-life. Others have a very brief half-life; strontium90, for example, has a half-life of 28 years. For researchers to determine the age of prehistoric materials, the halflife of an element must be of an appropriate period to determine the age of the material. For example, material composed of an element with a brief half-life would be gone in just a few years; none would remain in very old material for the purpose of dating. Carbon is the most useful element for isotopic dating (Figure 4.57). Carbon is present in all living things. It has several stable isotopes, including 12 C and 13 C, and the critical radioactive isotope, 14C, also known as carbon-14 or radiocarbon, with a halflife of approximately 5730 years. Because of this rather short half-life, materials older than about 40,000 years do not contain sufficient remaining
radiocarbon to be dated using this method. Carbon-14 is a very rare commodity; only 6 kg (13 lb) are produced each year in the atmosphere. There are only about 60 tons on Earth, so trying to find a carbon-14 atom in archaeological materials is a bit like trying to find a specific piece of gravel in a full dump truck. Carbon-14 is produced in the atmosphere by cosmic radiation, which was initially assumed to have been at as constant a level in the past as it is today. The radioactive carbon combines with oxygen, forming carbon dioxide, and is incorporated into plants in the same ratio in which it is found in the atmosphere. Animals eat those plants or other animals. When a plant or an animal dies, the intake of carbon ceases. Thus, the amount of radiocarbon in prehistoric material is a direct function of the length of time the organism has been dead. A variety of organic materials can be assayed by radiocarbon dating, including wood, bone, shell, charcoal, and antler. Carbon-14 often survives best at prehistoric sites in the form of charcoal, and this material has been most commonly dated using the radiocarbon method. However, wood charcoal can come from very old trees or wood and may not date the actual archaeological material accurately. If charcoal from the older rings of the tree is used for dating,
(1) Cosmic ray
14 N 14 N
14 C 14 C
atom
+
O
atom
O
=
Oxygen molecule (O2)
Proton
O
O
14 C 14 CO 2 molecule
(2)
(3)
Electron
Figure 4.57 The principles of radiocarbon dating. (1) Neutrons in cosmic rays strike nitrogen-14 atoms in the atmosphere, releasing a proton and creating carbon-14. (2) Carbon-14 and oxygen enter living organisms. (3) The death of an organism stops intake of new carbon-14 and decay begins. (4) Carbon-14 continues to disintegrate at a known, measurable rate.
(4)
In a 2-g sample: At death, there are 28 disintegrations per minute; 5730 years (one half-life) after death, 14 disintegrations per minute; 11,460 years (two half-lives) after death, 7 disintegrations per minute.
_ Concept Radiocarbon Dating
145
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Figure 4.58 The accelerator mass spectrometer (AMS) counts the number of radioactive atoms of carbon (14C) in a sample to determine how much is left and thus the age of the sample. The sample is disintegrated into individual atoms at the ion source and the atoms are sent through the instrument along the beam line. Atoms are then accelerated to separate them by weight. Both 12C and 14C are measured by instruments along the beam.
Page 146
Beam-bending magnet 14 C 14 C
measured Accelerator
detector
Beam-bending magnet
Magnetic lens Velocity filter High-voltage source Transformer
Beam line Ion source
Charcoal
Grain
Wood
Parchment
Bone
Charred bone
Figure 4.59 The size of samples needed for AMS dating is quite small, shown here compared to a U.S. penny.
radiocarbon dating An absolute dating technique based on the principle of decay of the radioactive isotope of carbon, 14C; used to date archaeological materials within the past 40,000 years. accelerator mass spectrometry (AMS) A method of radiocarbon dat-
ing using an accelerator to count the individual isotopes of the carbon sample.
_ _
calibrated dates Dates resulting from the process of calibration, the correction of radiocarbon years to calendar years, by means of a curve or formula derived from the comparison of radiocarbon dates and tree rings from the bristlecone pine. Calibration extends approximately 6000 years into the past.
146
Chapter 4 The Hunters
the age may be off by several hundred years. For more reliable dates, plants other than trees should be used. Materials with a short life, such as nutshells, corncobs, and small twigs, are preferred over wood charcoal. The actual measurement of radiocarbon is straightforward. A sample of known weight is cleaned carefully and burned to create a pure gas of carbon dioxide. The radioactive carbon isotopes in that gas are then counted. A Geiger counter is used to record radioactive emissions from the gas. Several grams of organic material are normally required to produce enough gas for counting. New technology, however, has resulted in a reduction in the amount of material that is measured. A device called an atomic accelerator mass spectrometer (AMS) is now being used to measure the ratio of carbon isotopes (Figure 4.58). Rather than measuring radioactivity with a Geiger counter, an AMS separates and counts individual carbon atoms by their weight, a much more accurate process. Moreover, less than 0.01 g of sample is needed, and individual pieces of charcoal or a single nutshell or cereal grain can be dated directly (Figure 4.59). Accelerator mass spectrometry is now being used to provide new dates for many sites and materials that could not previously be measured. The technique, which was implemented in 1980, has already produced very important results. There are certain problems with radiocarbon dating, some involving contamination of the sample. However, as the number of actual measurements in-
creased, it became clear to investigators that there were often regular errors in the dates. Because of those concerns, a detailed study of the method was undertaken to assess the error factor. The rings of known age from the bristlecone pine and other trees, such as the Irish oak, were compared against radiocarbon measurements of the same material. The bristlecone pine is one of the oldest living organisms on earth, reaching ages of up to 4000 years. Individual growth rings from the tree could be counted from the outer bark to the inner core, and the exact age of a ring could be measured by dendrochronology in calendar years (see “Chaco Canyon,” Chapter 7, p. 306). That same ring could also be dated by radiocarbon to determine the relationship between calendar years and radiocarbon years. This study recognized changes in the amount of radioactive carbon in the atmosphere over time and the need for a correction, or calibration, factor for radiocarbon dates to convert them to true calendar years. The assumption that radiocarbon has always been produced at a constant rate in the atmosphere was incorrect, and differences between radiocarbon years and calendar years are related to changes in the rate of cosmic ray bombardment. Radiocarbon dates that have been correctly converted to calendar years are called calibrated dates. The dates used for the past 10,000 years in this book are calibrated 14C dates.
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 147
Site
:
Beringia The arrival of the first inhabitants in the New World
The discoverers of North America walked to their new homeland sometime before 15,000 years ago, although the exact date is not yet known. These individuals were neither Irish monks in leather canoes, nor Scandinavian Vikings in longboats, nor Italians in tiny ships, but small groups from Asia who crossed, by foot, a land bridge connecting the two continents. The ancestors of the American Indians were the first inhabitants of the Western Hemisphere, a fact that was readily apparent when the “New World” was “discovered” by Columbus. Siberia and Alaska today are separated by less than 100 km (65 mi) at their closest point across the Bering Strait (Figure 4.60). The water in this strait is relatively shallow; the seafloor is no more than 50 m (160 ft) beneath the sea. During the colder intervals of the Pleistocene, when global sea levels were as much as 125 m (400 ft) below present levels, the floor of the Bering Strait became dry land. During periods of maximum cold and low sea level, this land area of Beringia, as it is called, would have been more than 1000 km (650 mi) wide and indistinguishable from the continents on either side. The warm Japanese Current swept the southern shore and kept most of the area free of ice. The region would have been relatively flat and treeless, a bleak and windswept plain. Scattered groups of mammoth (Figure 4.61), bison, horse, reindeer, camel, and many other species moved across this region during the cold periods of the Pleistocene. At some point, people followed these herds from northeastern Asia. There is no doubt that the first inhabitants of this New World came from Asia; the major question is when they came. By 8000 B.C., human groups had occupied most of both continents, from
Alaska to Tierra del Fuego at the southern tip of South America. The best place to begin considering the question of when this colonization started is at the point of entry. There were two possible barriers to the movement of terrestrial animals from northeastern Asia into central North America: (1) the Bering Strait, the body of water that separates the two continents today, and (2) an immense sheet of glacial ice that covered much of northern North America during the late Pleistocene. Beringia was exposed as dry land for long periods during the later Pleistocene and submerged only briefly
Figure 4.60 The New World and Beringia, the land bridge between northeastern Asia and Alaska, at the end of the Pleistocene, 12,000 years ago. The map shows the location of Beringia, the extent of continental glaciation, the coastline during lower sea level, and the sites mentioned in the text. These sites were occupied after the ice sheets had retreated. The dashed arrows show the possible route through a gap in the Canadian ice sheets.
ARCTIC OCEAN BERINGIA
GLAC IAL
IC E
Ushki Kennewick
Hebior/ Schaefer
Vail
Lindenmeier Cactus Hill
PACIFIC OCEAN
Monte Verde
_ Site Beringia
147
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 148
Figure 4.61 A skeleton of a woolly mammoth. Woolly mammoths roamed the area of Beringia and much of North America at the close of the Pleistocene. This extinct form of elephant is 4 m (13 feet) at the shoulder, about the height of the top of a basketball backboard.
Figure 4.62 Clovis Paleoindian spearpoints.
_ _
148
Chapter 4 The Hunters
during the warmer interglacials. However, for people to cross, they would first have to be present in northeastern Asia. An important set of clues was found on the Kamchatka Peninsula in eastern Siberia, which was part of Beringia during periods of lower sea level. Ushki I, excavated by Nikolai Dikov of the Northeastern Interdisciplinary Research Institute in Siberia, contained a grave filled with red ochre and several large, peanut-shaped dwellings. The artifacts included small, stemmed, bifacial flint projectile points similar to ones known from Alaska. The animal bones at the site revealed that the inhabitants hunted bison, reindeer, and probably mammoth, as well as catching large salmon. The lowest levels at Ushki I date to around 14,000 years ago and provide evidence that people making bifacial artifacts (a technology employed in the production of Paleoindian points such as those shown in Figure 4.62) were present in this area at least 4000 years before the end of the Pleistocene. Although the exact timing is debated, Beringia was probably dry land for long periods between 60,000 and 13,000 years ago. After 13,000 B.P., warmer climatic conditions began to melt the continental ice sheets, and the meltwater gradually returned the seas to their present levels. During the period of maximum cold and lower sea level, the continental ice sheets of northern North America stretched from the Aleutian Islands across Canada to Greenland. Alaska was cut off from the rest of North America and was more a part of Asia than of the New World. A gap or corridor between the western and eastern centers of the Canadian ice sheets was probably open between 40,000 and 21,000 years ago and again shortly after 14,000 B.P. This corridor, when open, may have provided the route for the penetration of these newly arrived hunting groups into North and South America. Another possible route for the spread from Alaska into the continent lay along the western coastline of North America. That rich coastal habitat would have been ice-free and likely
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Figure 4.63 The reconstructed haft and mounting of a Paleoindian point. Note how the shaft fits neatly into the flute on the spearpoint.
accessible to the hunters crossing the Bering Land Bridge. It is also possible that the strait was crossed at other times on the sea ice or even by boat. Whatever the route, these emigrants must have been few in number, because they left little evidence of their presence. Sites from this early period are rare, and the dates are usually debated. A series of new datings and discoveries, however, have pushed back the record for the early inhabitants of the Americas. Monte Verde in Chile, dating to at least 13,000 years ago, provides good evidence for these early inhabitants. Monte Verde contains the remains of extinct elephants and other animals in association with stone tools. At the site of Cactus Hill, Virginia, the inhabitants built hearths and left stone tools that may be as old as 15,000
Page 149
years. Finds of human remains from Mexico and elsewhere have been dated to 13,000 years ago. Mammoths in southwestern Wisconsin bogs have been found associated with spearpoints and dated to as much as 13,500 years ago. Underwater exploration off the west coast of British Columbia, Canada, has revealed human artifacts at a depth of 50 m below sea level, suggesting a very early date, as well as supporting the idea of a coastal route for the first inhabitants. The evidence for the arrival of the first Americans before 12,000 years ago is growing rapidly. Perhaps the most remarkable recent discovery regarding the first Americans comes from a series of small caves in southern Oregon. Evidence for the presence of humans on the continent is vivid and well preserved in the dry layers of the Paisley Caves. The indications of human activity in the caves comes from camel and horse bones with cutmarks, a few questionable stone tools, and a dozen or more coprolites. In the dry interior of the caves, these objects have not changed much over time and still resemble human feces. To be certain that these turds were human, the ancient DNA preserved in them was examined and their origin confirmed. Because the feces were associated with extinct fossil species and should be old, the archaeologist in charge of the excavations, Dennis Jenkins, got a series of radiocarbon dates directly from the coprolites. These dates were very surprising, falling around 14,250 years ago, the oldest reliable dates for humans in the New World. Now a date of at least 15,000 years ago for the colonization of the Americas looks more and more reasonable. Slightly later archaeological remains in North America, from the beginning of the Holocene, are generally known as Paleoindian. Sites from this period are recognized by the presence of a specific type of stone spearpoint, known as a fluted point (Figure 4.63). Paleoindian spearpoints document the presence of early Americans in North America between 11,000 and 9,000 years ago.
Paleoindian The period of largegame hunters in North America at the end of the Pleistocene. fluted point The characteristic arti-
fact of the Paleoindian period in North America. The flute refers to a large channel flake removed from both sides of the base of the point to facilitate hafting.
Site Beringia
149
_
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 150
Figure 4.64 The distribution of Clovis points in the United States, showing a majority east of the Mississippi. The distribution probably reflects modern population density as well as a prehistoric pattern. More points are usually found where more people live.
200 100 75 50 25 10 5 2 0
Clovis An archaeological culture dur-
ing the Paleoindian period in North America, defined by a distinctive type of fluted point. Folsom An archaeological culture
_ _
during the Paleoindian period in North America, defined by a distinctive type of fluted point and found primarily in the Great Plains.
150
Chapter 4 The Hunters
The best-known points from this period are called Clovis, after the original find spot in eastern New Mexico. Clovis people were relatively mobile groups of big-game hunters, spreading across most of North America east of the Rocky Mountains. A series of new radiocarbon dates from Clovis sites across North America suggests that this archaeological culture was a short-lived phenomenon, lasting approximately 200 years. This new information indicates that Clovis was the result of a very rapid spread of new technology and perhaps new people across the continent. It is remarkable that such a brief archaeological episode has left such lasting traces, but the distinctive Clovis spearpoints are a vivid and visible signal of that time period (Figure 4.64). New types of projectile points evolved following the Clovis period. Distinctive, regional types of Paleoindian points are found throughout the New World. Folsom points, for example, are found only in the Great Plains. A number of sites are known where animals were killed and butchered with these points, but only a few actual settlements have ever been found. The Lindenmeier site in northern Colorado is one of the rare examples of a Folsom campsite.
In sum, it appears that Homo sapiens sapiens came to the New World for the first time before 15,000 years ago. Groups of hunters quickly expanded across both continents and were likely involved in the extinction of a number of animal species (see “Pleistocene Extinction,” p. 151). It is important to remember that the Bering Strait has probably been the point of entry for several expansions into the New World. The colonization of North and South America was not a single event but a series of crossings. Certainly, the ancestors of modern Eskimos came across the strait relatively late, perhaps 4000 years ago, in boats. Thus, the timing of the arrival of the first Americans is a difficult and complex issue that will continue to be investigated.
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 151
Concept
<
Pleistocene Extinction Natural versus human causes
Discoveries of the remains of very large animals that once roamed North America in various places—the La Brea tar pits in Los Angeles, the caves of the Grand Canyon, the bone beds of eastern Missouri—have attracted a great deal of attention. Many of these animals were large carnivores and herbivores that wandered the continent over the past 2 million years. By the end of the Pleistocene, however, some 35 species of land mammals, nearly half the total number, had become extinct in North America (Figure 4.65). The giant sloth and giant beaver, the horse, camel, mammoth and mastodon, lion, cheetah, and short-faced bear, all much larger than their modern counterparts, had become extinct by 10,000 years ago. The giant sloth was about the size of a giraffe and weighed up to 3 tons. Caves in the Grand Canyon have preserved many examples of softball-size giant sloth
Ovis
Euceratherium
Equus asinus
dung. The giant beaver in the Great Lakes area weighed as much as 140 kg (300 lb). A similar, though not as complete, pattern of extinction in large mammals occurred at the close of the Pleistocene in Europe and Asia, where the mammoth, woolly rhino, cave bear, lion, and other species also became extinct. What caused the demise of so many animals in such a relatively short period of time? Scientists, archaeologists, paleontologists, climatologists, and others have considered that question for decades. Two major explanations have been proposed: climatic change (a natural cause) and hunting overkill (a human cause). Each side in the debate has a set of facts and conjectures that appears to contradict that of the other. For example, those supporting overkill argue that very similar climatic changes during earlier interglacials did not result
Bison
Odocoileus
Mammuthus
Navahoceros
Stockoceros
Camelops
Figure 4.65 Some Pleistocene and early Holocene animals of North America now extinct. These largely extinct species are related to elephants, camels, horses, and other creatures that formerly roamed the continent.
Equus
Antilocapra
_ Concept Pleistocene Extinction
151
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 152
Figure 4.66 The neck vertebra of a bison with a Paleoindian point embedded in it (at arrow).
42%
Mollusks
41%
Small mammals
50%
_ _
Figure 4.67 Percentages of extinct animals compared with the total number of species in each group.
152
Chapter 4 The Hunters
Large mammals
in the extinction of many species, such as the mammoth, which were adapted to a very broad range of environments. Also, the widespread appearance of Clovis hunters around 11,000 years ago coincides closely with the demise of several of the extinct species. Fluted points have been found in association with the bones of extinct mammoth, mastodon, horse, tapir, camel, and bison (Figure 4.66), suggesting that Paleoindian hunters, with easy access to animals virtually ignorant of human predators, quickly eliminated entire species of large animals they encountered. Those who doubt the role of human hunters in the disappearance point out that certain extinct species are never found at kill sites. There is, in fact, no direct evidence of human predation on animals such as the giant sloth and beaver. Nonmammalian species also suffered extinction at the end of the Pleistocene (Figure 4.67). For example, almost 45% of the bird species in Pleistocene North America had disappeared by 11,000 years ago. There is little reason to suspect human intervention in the extinction of those species. New evidence for increasing aridity at the end of the Pleistocene and its effect on animal populations argues against the role of human hunters as the sole factor in the extinctions. This climatic change led to a shift in the distribution of many species of plants and animals and led to the extinction of large animals. Requiring more food and more space to live, large animals, as a group, are more vulnerable to changes in their environment. They are also found in lower numbers than small animals, so the loss of a few individuals in an area could have serious consequences during mating season. Larger animals, too, have a longer generation length than smaller ones, so that genetic changes move more slowly through the population. This also makes larger animals like the mammoth more susceptible to extinction (Figure 4.68). Proponents of the climate change theory argue that seasonal differences in climate were not as pronounced in the Pleistocene as they are today. Late
pri31057_ch04.pp104-163.qxd
Figure 4.68 China.
10/1/09
7:55 AM
Page 153
A mammoth skeleton on display at the National Natural Museum, Beijing,
Pleistocene winters must have been warmer, and summers cooler, than they are at present. As seasonal differences increased at the close of the Pleistocene, environmental changes would have been dramatic. Such difficulties, in conjunction with effective human predation, may have been too much for those species of large animals that died out. A very recent theory ignores climate change and human predation in favor of meteor impact. James Kennett
and his colleagues point to the presence of “nanodiamonds” in a layer found in Europe and North America dating to ca. 12,900 years ago. They suggest that comets exploded over the North American ice sheet, causing catastrophic melting and a cold episode that may have wiped out mammoths and other large species and brought an abrupt end to the Clovis culture. This highly controversial idea is going to require some time to fully evaluate.
_ Concept Pleistocene Extinction
153
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 154
Site
:
Monte Verde Early hunter-gatherers in South America
Figure 4.69 Monte Verde, 1983: a general view of the site and field laboratory. Excavation trenches can be seen running perpendicular to the creek in the lower right of the photograph. The buildings in the center of the photograph are part of the excavation headquarters.
_ _
154
Chapter 4 The Hunters
Traces of human groups in the Americas before 12,000 years ago are almost nonexistent. There are only a few examples of archaeological sites in the New World that contain both definite evidence of an early human presence and reliable radiocarbon dates. In many instances, sites that were candidates for very early human occupation have
been discounted because of contaminated carbon samples or questionable stone artifacts. The best examples of human occupation sites before 10,000 B.C. are in South America, rather than nearer the original point of entry across the Bering Strait. Monte Verde (MON-tay VERday) is a 13,000-year-old residential site in the cool, forested region of northern Chile (Figure 4.69). The site was discovered in 1976 by Tom Dillehay, of the University of Kentucky. While surveying in the area, he discovered bone and stone artifacts in a shallow, swampy area—a peat bog—along Chinchihuapi Creek, a small, slow stream that drains this part of the rain forest of southern Chile. The site lies along the sandy banks on either side of the creek (Figure 4.70). Excavations since the original discovery have uncovered a number of remarkable and unexpected finds for such an early site. Excellent conditions of preservation resulted in the recovery of plant remains and numerous wooden objects, along with stone flakes and broken animal bones. Wood is rare at most archaeological sites, but it was preserved in the bog at Monte Verde. Apparently, the bog developed during or shortly after the abandonment of the site and quickly enclosed all the remaining materials in a mantle of peat. Peat provides a waterlogged, oxygenfree environment where such objects can be preserved. The timber and earthen foundations of perhaps 12 living structures were recovered in excavations (Figure 4.71). The rectangular foundations, made of logs and planks held in place with stakes of a different kind of wood, enclose rooms 3–4 m (10–13 ft) long on each side. Posts were placed along the foundation timbers and supported a framework of saplings, which may have
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 155
0
Upper terrace
20 m
Shallow bogs and marshes
Foundations
Foundations
Creek in13,000 B.P.
Lower terrace
Sandbar Modern creek Foundations Lower terrace Upper terrace
been covered with animal skins. Small pieces of what may be animal hide were preserved next to the timber foundation. Two large hearths and a number of shallow clay basins provided fireplaces for the inhabitants of these huts. Even a child’s footprint was found preserved in the hardened clay of the basin of one of the small fireplaces. Many of the artifacts at the site were found inside the structures. Wooden artifacts include digging sticks, tool handles, spears, and a mortar or basin, in addition to the material used for construction. Several kinds of stone tools, both flaked and ground, were found. No stone projectile points were made here, but evidence of other weapons was uncovered. Spherical stones with an encircling groove were probably bola stones, a South American throwing weapon with three leather thongs weighted at each end. The bola is thrown in a spinning fashion, and the stone weights wrap the thongs around the prey. Other stone balls without grooves were likely used in a sling as heavy projectiles.
To the west of the living floors was a single, more substantial, rounded structure with a pointed end and a foundation of sand and gravel. The bases of wooden posts at the edge of the foundation mark the walls of the structure; a kind of yard or patio was marked off with branches near the entrance. Mastodon bones, animal skins, stone and wooden tools, salt, and the remains of several types of plants were found in this area. The arrangement of these structures over an area of roughly 70 × 100 m (220 × 400 ft, about the size of a soccer field) suggests a well-organized community. Dillehay estimates that the site was occupied for perhaps a year by 20–30 people. The occupants of Monte Verde apparently relied primarily on plants and large animals for their livelihood. Most of the bones came from mastodons (a close but extinct relative of the elephant and the mammoth), a type of llama, amphibians, reptiles, and birds. Plant remains at the site are from species that ripen throughout the year, suggesting that this was a
Figure 4.70 The location and plan of the Monte Verde site. The settlement lay on both sides of the creek, covering an area of approximately 7000 m2, as indicated by the dashed line. Foundations are outlined where they have been located.
_ Site Monte Verde
155
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 156
Figure 4.71 Excavated structures at Monte Verde. The photo shows the remains of wooden hut foundations.
y
Why did it take so long for our ancestors to reach the New World?
_ _
156
Chapter 4 The Hunters
year-round settlement. The wetlands (marshes, bogs, and streams) of the Monte Verde area are a very rich environment, and many edible plants grow there today. The plant remains document extensive gathering of the local vegetation. A total of 42 edible species of plants have been identified from the site. Most of the evidence is from tubers and roots that were preserved, including the wild potato. This is the oldest evidence anywhere for the potato, which was later domesticated in the Andes Mountains. The products of other plants, including seeds, berries, nuts, and fruits, were also recovered during the excavations. The number of grinding stones at the site points to the importance of plants in the diet at Monte Verde. Herbaceous plants were found, which today are used for medicinal, rather than nutritional, purposes. A chewed cud from two kinds of seaweed and the leaf of a tree was perhaps medicinal. Exotic objects, including several of the plant species, beach-rolled pebbles, quartz, and bitumen (an adhesive tar), were also brought to the site from the Pacific coast about 25 km (15 mi) to the south. The evidence from Monte Verde contradicts most traditional views of
the peopling of the Americas. The radiocarbon date of 11,000 B.C. for this site in South America documents the fact that early Asians crossed the Bering Strait before 13,000 years ago. The information from Monte Verde has forced a reconsideration of our interpretation of the earliest inhabitants of the New World, because it is in direct contrast to what has been observed at most sites where preservation is not nearly as good. Paleoindian sites normally contain small concentrations of stone artifacts, sometimes in association with the bones of large, extinct animals. Paleoindians are thought to have lived primarily as small, mobile groups of big-game hunters. The evidence for permanent residence at Monte Verde is in direct contrast to Paleoindian occupations elsewhere. The organic materials indicate the importance of plants, as well as animals, in the diet at that time. The existence of wood and wooden tools, more common at Monte Verde than stone artifacts, provides an intriguing look at the organic component of tools and equipment rarely seen in the archaeological record.
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 157
Site
:
Lindenmeier Late Pleistocene hunters in Colorado
Lindenmeier, located in northern Colorado at an elevation of 2000 m (6500 ft) in the foothills of the Rocky Mountains, lies at the intersection of three major environmental zones: the eastern slope of the Rockies, the Colorado Piedmont, and the High Plains (Figure 4.72). Gullies cutting through sedimentary deposits in this area washed out and exposed the buried bones of large bison and large spearpoints. The local ranchers, who discovered these remains in 1924, wrote to the Smithsonian Institution in Washington, DC, to inform them of the site’s existence. In 1934, Frank Roberts, of the Smithsonian, was sent to investigate the site. Roberts wanted to document the association of hunters and extinct animals in the past. Early sites in North America were almost unknown at the time. Roberts’ excavations at Lindenmeier continued each year from 1934 through 1940. The field crew spent almost 600 days on the excavations, opening over 1800 sq m (almost half an acre) in the process and digging deeply into the buried archaeological deposits (Figure 4.73). This project eventually uncovered one of the largest Paleo-
indian sites in the New World. Roberts’ careful excavations and recording procedures allowed the final results of his original investigations to be published by others after his death. Radiocarbon dates from the site place the occupation of Lindenmeier at 9000 B.C., just before the end of the Pleistocene. The climate at that time was less arid and cooler than it is today. The local environment at the site was more wooded, with stands of juniper and pine. Fragments of charcoal from these species were recovered from fireplaces at Lindenmeier. Ground cover would have been heavier, with thick grasses over much of the area. The area around the site can be envisioned as a lush meadow watered by an active spring, which attracted both animals and their predators. The spring, adjacent to the site, must have been the reason this spot was selected as a campsite. More than 15,000 animal bones were recovered in the excavations. Twelve species were represented at Lindenmeier, including wolf, coyote, fox, hare, rabbit, turtle, deer, antelope, and bison; they inhabited all three
Figure 4.72 The site of Lindenmeier (at the juncture of the arrows) located between the plains and the piedmont in eastern Colorado.
_ Site Lindenmeier
157
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 158
Figure 4.73 The 1937 Smithsonian Institution excavations at the site of Lindenmeier.
environments near the site. Antelope, wolf, and fox are native to the Rocky Mountains and the piedmont; bison, coyote, and jackrabbit are common in the High Plains. The bison was by far the most common animal represented at the site and must have provided most of the meat consumed there. This was a now-extinct form of bison, a huge animal some 2 m (6.5 ft) at the shoulder. In all likelihood, the Folsom hunters cooperated to drive these animals into blind canyons and other traps where they could more easily be killed.
Figure 4.74 A Folsom Paleoindian point.
pressure flaking A technique for
_ _
producing stone artifacts by removing flakes from a stone core by pressing with a pointed implement.
158
Chapter 4 The Hunters
Figure 4.75 A Paleoindian flintknapper removing the flute from a Paleoindian point. A chest crutch is used to apply pressure to remove the flake from the base of the point. This flute, or channel, facilitates hafting.
More than 50,000 stone artifacts have been counted from the excavated areas at Lindenmeier. Approximately 5000 of these stone pieces and 70 bones were shaped into finished tools. The bone tools included needles and simple pointed pieces. In addition, a number of bone pieces decorated with notches and engraved lines, and several bone beads, were excavated. More than 600 projectile points were recovered, including almost 250 Folsom points (Figure 4.74). These Paleoindian spear tips are slender, bifacially worked stone points, shaped carefully on all surfaces by pressure flaking. To finish the point, a single long flake, or flute, is removed from the base of each side as a channel to facilitate hafting to a wooden spear shaft. This fluting flake was probably removed by pressure using a chest crutch and vise (Figure 4.75). Paleoindian points are often found with the skeletons of big game animals such as giant bison, mastodon, and mammoth. Often broken points were either resharpened or reworked into other tools, such as scrapers and knives. The repair of broken equipment was one of the characteristic activities at these campsites. Fifteen distinct concentrations of archaeological materials were observed in the excavations at Lindenmeier. Two groups of these concentrations, designated Area I and Area II, deserve special attention. Were these two collections of artifacts and bones left by the same group on separate visits or by two distinct groups, perhaps there at the
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
same time? Careful measurements of the Folsom points in the two concentrations indicated two sizes. The points in one area were, on average, slightly smaller than, and fashioned in a different manner from, those in the other area. These differences suggest that two different groups of people were responsible for the two concentrations. Further evidence for the differences between the two areas comes from the original sources for the obsidian found at the site. Obsidian is a
Page 159
type of natural glass produced by volcanic eruptions and is highly prized for making stone tools. Sources for obsidian are limited in number and size. The obsidian in Area I came from New Mexico, and the material in Area II came from the north, near Yellowstone National Park in Wyoming. Lindenmeier thus seems to provide evidence of two or more different social groups that came together, perhaps to cooperate in bison hunting, for a brief time some 11,000 years ago.
obsidian Translucent, gray-to-black or
green, glasslike rock from molten sand.
_ Site Lindenmeier
159
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 160
Site
:
Kennewick Man An early American
Figure 4.76 The Kennewick skull, and the artistic reconstruction of the face.
160
Chapter 4 The Hunters
In recent decades, our picture of the first inhabitants of the New World has changed dramatically because of the discovery of a number of new sites and skeletons. One of the most important of these finds is the human skeleton found near Kennewick in the state of Washington. The Kennewick material has provided major lessons in both archaeology and human relations. Spectators at a boat race on the Columbia River in 1996 came across a human skull washing out close to the riverbank near the town of Kennewick, Washington. Over the next month, many other bones were found, and by the end of August that year, a radiocarbon date of 7500 B.C. revealed that Kennewick was one of the earliest human skeletons in the New World. The discovery of Kennewick caused a major controversy about the ownership of the remains. Native American groups in the region wanted to rebury an ancient ancestor. Scientists wanted to study this early and unusual skeleton. This important matter was in the courts for years, and eventually, it was decided to let the scientists study Kennewick man in detail. The ethical debate surrounding this human skeleton is discussed in more detail in Chapter 12. The original study of the Kennewick remains led the investigator, James Chatters, to conclude that they were probably from a more recent European descent because of the very
good preservation of the bone, an absence of distinctive Native American features, and the location of an early European homestead near the site (Figure 4.76). However, Chatters subsequently noticed a stone projectile point, embedded in the pelvis of this individual, of a type known to be at least 4500 years old in the Pacific Northwest of North America. Further examination of the Kennewick man revealed that his body had been intentionally placed in a grave on his back with his arms at his sides and his palms down. The skeleton was very well preserved, and all the teeth were intact in the jaw. The man was somewhat older, age 40–55, slender, and tall (around 173 cm; 5 ft, 7 in). He had endured a number of injuries in his lifetime, including a compound fracture of the rib cage and significant damage to his left shoulder, in addition to the projectile point in his pelvis. The cause of death could not be determined since all the observed injuries had healed. Analysis of amino acids and carbon and nitrogen isotopes in the bone revealed that fish was a substantial component of his diet. This skeleton of a male individual is important not only because of its antiquity but also because of the unusual nature of the remains. The questions raised by the unusual features of the skull demand further investigation to better understand the original inhabitants of the Americas.
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 161
Images and Ideas The End of the Paleolithic
T
he story of the Paleolithic is a remarkable saga—an evolutionary journey from primate to human. The major changes that occurred in human biology and culture made us essentially what we are today. We began more than 4 million years ago as chimpanzee-like primates living in open grassland environments of subtropical Africa. The climate was mild, plants grew year-round, and large predators killed and ate many animals, leaving behind bits of meat and bone marrow for hungry hominins. Sharpened edges of stone helped remove the meat, and other stones cracked the heavy bones. Biologically—lacking claws, big teeth, and speed—early humans were ill-equipped to defend themselves or their young from the predators of the plains. As social beings, however, with safety in numbers, they may have been able to drive away those ferocious carnivores. The early hominin adaptation was a successful one that spread over most of Africa by about 2 million years ago. Brain size almost doubled during this period, and faces underwent dramatic changes. The handaxe was invented—a marvelous, multipurpose tool sculpted from a slab of stone. As hominin numbers and adaptive success increased, early humans began to move out of the African cradle into Eurasia, where a cooler, more temperate climate challenged ingenuity. The mysterious force of fire was controlled during this time for heat, light, and cooking, becoming a new ally in the continuing battle with nature. Plants were not available during the northern winters, and survival came to depend more on the hominins’ abilities as hunters. By 100,000 years ago, our ancestors had occupied much of the Old World, including the cold tundras of Pleistocene Europe and Asia. The human brain reached modern size. New tools and ideas prevailed against the harsh environment. Life became something more than eating, sleeping, and reproducing. Burial of the dead and care of the handicapped and injured illustrate a concern for fellow hominins. A cultlike preoccupation with cave bears and some evidence for cannibalism suggest a concern with the supernatural; ritual and ceremony achieved a place in hominin activities. The Upper Paleolithic was the culmination of many trends—in biology and culture, in language and communication, in ritual and ideology, in social organization, in art and design, in settlement and technology—that had begun several million years earlier. Evolution brought humanity to our modern form, Homo sapiens sapiens. New continents were explored; Australia, North America, and South America were colonized. More kinds of implements were made from a wider variety of materials than ever before. Bows, boats, buttons, fishhooks, lamps, needles, nets, spearthrowers, and many other items were produced for the first time during this period. The dog was domesticated as a faithful hunting companion and occasional source of food. Caves and many artifacts were decorated with paintings, carvings, and engravings, as an awareness of art and design developed in the human consciousness (Figure 4.77). Sites from the Upper Paleolithic were larger and more common than those from previous periods. From almost any perspective, this period of the Upper Paleolithic represents a dramatic change in human behavior, almost certainly associated with changes in the organization of the brain or in the use of language, or both. Essentially modern behavior appeared following this transformation, and the rapid change from hominin to human, from archaic to modern, from the past to the present had begun. The end of the Paleolithic was likely the apogee of hunter-gatherer adaptations. Successful groups of foragers lived and increased in almost all the environments on Earth. It was, in fact, this expansion in numbers that was partly
< ( : < V : y < p : W
Figure 4.77 A highly romanticized 1870 engraving of the noble savages of the Paleolithic.
Images and Ideas The End of the Paleolithic
161
_
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 162
responsible for the end of a hunter-gatherer way of life. Increasing populations required new and more productive sources of foods. The bounty of the land, the wild plants and animals of nature, simply were not enough to feed everyone. Experiments to increase the available amount of food were necessary. The story of the domestication of plants and animals and the beginning of the Neolithic is the subject of Chapter 6. Chapter 5 describes events after the end of the Pleistocene, leading to the beginnings of agriculture.
p
DISCUSSION QUESTIONS 1. What caused the disappearance of Neanderthals in western Eurasia?
w w w. m h h e . c o m / p r i c e i p 6 e
For more review material and study questions, see the self-quizzes on your online learning center.
2. The Upper Paleolithic has been described as a time when human creativity expanded enormously. What kinds of evidence document these changes? 3. What are the major changes that mark the end of the Pleistocene and the Paleolithic? 4. Where did the first fully modern humans come from, and where did they go? 5. The arrival of the first Americans is one of the more controversial subjects in archaeology today. What is your impression, based on the available evidence?
z
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e.
Bahn, P., and J. Vertut. 1997. Journey through the Ice Age: Art and architecture. Berkeley and Los Angeles: University of California Press. Breathtaking photographs of the creative explosion. Bonnichsen, R., ed. 2004. Who were the first Americans? Corvallis, OR: Center for the Study of the First Americans. The latest evidence from the artifactual and skeletal remains of the original inhabitants of the Americas. Clottes, J. 2008. Cave art. London: Phaidon Press. A new view on the making of cave paintings. Fagan, B. 1987. The great journey. London: Thames & Hudson. A very readable account of the arrival of the first Americans. Findlayson, C. 2004. Neanderthals and modern humans: An ecological and evolutionary perspective. Cambridge: Cambridge University Press. An up-to-date review of the evidence for Neanderthal extinction and the role of modern humans. Harvati, K., and T. Harrison, eds. 2007. Neanderthals revisited: New approaches and perspectives. New York: Springer. Hoffecker, J. F., W. R. Powers, and T. Goebel. 1993. The colonization of Beringia and the peopling of the New World. Science 259:46–53. A technical synthesis of the evidence from northern North America and Asia. Klein, R. 2002. The dawn of human culture. New York: Wiley. A new view of what it means to be fully human. Lewis-Williams, D. 2002. The mind in the cave: Consciousness and the origins of art. London: Thames & Hudson. _ _
162
Chapter 4 The Hunters
A recent consideration of the explanations for cave art with new insights from the ethnographic record. Lieberman, P. 1991. Uniquely human: The evolution of speech, thought, and selfless behavior. Cambridge, MA: Harvard University Press. The evolution of language and other very human behaviors. Meltzer, D. J. 2009. First peoples in a new world: Colonizing Ice Age America. Berkeley: University of California Press. The latest evidence on the colonization of North and South America. Mulvaney, J., and J. Kamminga. 1999. The prehistory of Australia. Washington, DC: Smithsonian Institution Press. The latest summary of Australian archaeology. Schrenk, F. 2008. The Neanderthals. London: Routledge. Tattersall, I. 1999. The last Neanderthal: The rise, success, and mysterious extinction of our closest human relatives. Boulder, CO: Westview Press. A good summary of the Neanderthal question. Thomas, D. H. 2000. Skull wars: Kennewick Man, archaeology, and the battle for Native American identity. New York: Basic Books. An examination of the Kennewick lawsuit and the ongoing conflict between Native American tribes and the scientific community over who owns the rights to archaeological finds. White, R. 1986. Dark caves, bright visions: Life in Ice Age Europe. New York: American Museum of Natural History. A wonderfully illustrated volume on the art, both cave paintings and portable objects, of the Upper Paleolithic.
pri31057_ch04.pp104-163.qxd
10/1/09
7:55 AM
Page 163
_
pri31057_ch05.pp164-197.qxd
10/1/09
C
H
7:59 AM
A
Page 164
P
5
T
E
R
F
I
V
_ _ Figure 5.1 Burial of mother and infant son from Mesolithic Denmark, 5000
B.C.
E
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 165
Postglacial Foragers Introduction The World after 8000 B . C . New solutions to changing environments
F
igure 5.1 shows a photograph of a double burial from the Stone Age cemetery at Vedbaek, north of Copenhagen, Denmark. The burial belongs to the Mesolithic period, dating to around 5000 B.C. There are a number of remarkable graves in this cemetery. This particular burial is of a mother and infant child. The 18-year-old mother’s head rests on a bundle of tooth pendants that would have decorated a hide blanket or a piece of clothing. The newborn was buried next to his mother with a stone knife at his waist, like all the males in the cemetery. The child’s body was placed on the wing of a swan. An artist’s reconstruction of this grave can be found on page 173. These remarkable graves also point to the rapid developments that were taking place in human society. The pace of change had increased dramatically over time. For example, there were more innovations in artifacts and behavior during the Upper Paleolithic than in all the preceding periods. A number of major changes also took place at the end of the Pleistocene, beginning around 11,000 years ago. Large-game hunting, an adaptation that had characterized human prehistory for much of the Middle and Upper Paleolithic, began to decline as certain large-animal species became extinct and environments changed in response to warming climatic conditions. Human diet became more diversified and included more plant and animal species. This chapter examines the early part of the Postglacial period. In a few areas—Southwest Asia, the Far East, and parts of the Americas—plant and animal domestication appeared in the early Postglacial and began to alter the longstanding hunting-and-gathering pattern of human subsistence. These origins of agriculture are discussed in Chapter 6. In other places, hunter-gatherers continued their way of life, adapting to the changing environmental conditions of the Holocene. The term hunter-gatherers refers to human groups (also known as foragers) who use only the wild, natural resources of the earth, hunting animals, fishing, and collecting plants, nuts, seeds, shellfish, and other foods for their sustenance. Although different terms are used in the Old World and the New World for this period—Archaic in the New World and Mesolithic in the Old World—the basic way of life was very similar. The term Mesolithic, or Middle Stone Age, designates the period between the end of the Pleistocene and the beginnings of agriculture in Europe, North Africa, and parts of Asia. Only in the past 30 years has the significance of the early Postglacial period been recognized. Since the beginning of the 1800s, the Upper
165
< ( : < V : y < p : W
CHAPTER OUTLINE
Introduction The World after 8000 B.C. 165 CONCEPT The Postglacial Environment of Europe 169 SITE Vedbaek
171
CONCEPT Bone Chemistry and Prehistoric Subsistence 176 SITE Elands Bay Cave and De Hangen 179 SITE Sannai Maruyama SITE Carrier Mills
187
CONCEPT The Human Skeleton 192 CONCEPT Contemporary Hunter-Gatherers 194
Images and Ideas The World of HunterGatherers 196
183
w w w. m h h e . c o m / p r i c e i p 6 e
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
_
pri31057_ch05.pp164-197.qxd
10/15/09
10:51 AM
Paleolithic and the Neolithic have been acknowledged as important episodes in human prehistory. The Mesolithic, however, was thought to have been a period of cultural degeneration, occurring between the time of the spectacular cave paintings of the late Paleolithic and that of the farming communities of the Neolithic. Today, however, the Mesolithic is recognized as a time of intensification in human activities and organization (Figure 5.2). In their adaptations, these hunter-gatherer societies were similar to other groups in Africa, eastern Asia, and North and South America, consuming a wide range of wild plant and animal species and using a highly specialized technology. An incredible range of fishing gear, including nets, weirs, hooks, and harpoons, was developed during this period. Ground stone artifacts appear as axes, celts, plant-processing equipment, and other tools. Projectile weapons were equipped with a variety of tips made of bone, wood, antler, or stone. In those areas of Europe where bone and other organic materials have been preserved, artifacts are often decorated with fine, geometric designs. Cemeteries that are sometimes present at Mesolithic sites suggest more sedentary occupations. An example of a Mesolithic settlement discussed in this chapter is found in Denmark, where excavations at Vedbaek have exposed a cemetery and settlement dating to about 5000 B.C. Information from this site provides a good picture of early Postglacial hunter-gatherers. In South and East Africa, the term Late Stone Age is used to refer to the artifacts and the encampments of Holocene hunter-gatherers. Although most studies focus on the earlier periods of human prehistory in Africa, new findings from Elands Bay Cave on the west coast of South Africa proThe Far Side® by Gary Larson © 1993 FarWorks, Inc. All Rights Reserved. Used with permission. vide some indication of the way of life of these groups shortly before the introduction of herding and farming. In Japan, the period between the beginning of the Holocene and the introThe similarity of Postglacial duction of rice cultivation is known as the Jomon. Thousands of settlements were adaptations in many parts of occupied during this time, and these food-collecting peoples created very sophisthe world is remarkable. What ticated pottery. In fact, some of the earliest ceramics in the world were made in is responsible for this pattern? Japan approximately 12,000 years ago. Sannai Maruyama in northern Japan is a remarkable example of a Jomon settlement around 5500 years ago. In North America, the Paleoindian period of big-game hunting ended approximately 9000 years ago, about the same time many species of big game became extinct. The period between 6000 and 1000 B.C. is known as the Archaic and is very similar to the Mesolithic in Eurasia. Human groups began to exploit hunter-gatherer A hunter of large a broad spectrum of food sources. Many new sorts of subsistence pursuits seem to wild animals and gatherer of wild have begun at this time. Ground stone tools such as mortars and grinding stones plants, seafood, and small animals, as have been found at some Archaic sites, indicating an increasing reliance on plant opposed to farmers and food producers. foods. Exotic materials in some regions document an increase in long-distance Mesolithic The period of time of hunter-gatherers in Europe, North Africa, trade for obsidian, copper, and shell. The date for the end of the Archaic varies in and parts of Asia between the end of different areas of North America. In some regions, such as the West Coast, the the Pleistocene and the introduction of basin and range landscapes of what is now the western United States, and the subfarming; the Middle Stone Age. arctic and arctic reaches of Canada and Alaska, hunting and gathering persisted up Archaic The term used for the early to and beyond European contact. In other areas, such as the southwestern United Holocene in the New World, from apStates (Arizona, Colorado, New Mexico, Utah) and the major river valleys of the proximately 6000 B.C. to 1500–1000 B.C. midwestern and southeastern United States, a more sedentary way of life involvcultigen A cultivated plant. ing the use of cultivated plants, or cultigens, began as early as 1000–500 B.C., and
y
_ _
Page 166
166
Chapter 5 Postglacial Foragers
e
a
op
Am
er ic
Eu r
No r
th
es te rn W
og ic a Cl le im po at ch ic s ph as es
eo l G
Warmer 1
SubAtlantic
2
FARMING
HOLOCENE
3 4 5
7
Subboreal
BRONZE AGE
NEOLITHIC Atlantic
6 Climatic optimum
ARCHAIC MESOLITHIC
8
Boreal
9 Preboreal
10 11 12 Cooler
PLEISTOCENE
perhaps even earlier. In fact, recent evidence from plant genetics suggests that the domesticated gourd was carried into the Americas by some of the early settlers almost 10,000 years ago. Great diversity in Archaic adaptations was also seen in eastern North America, with major emphasis on fishing, hunting, and plant and nut collecting. Archaic sites such as Carrier Mills in southern Illinois document adaptations typical of this area. Settlements were often located along lakes, rivers, and coastlines to take advantage of aquatic resources. Piles of freshwater mussel shells, the remains of prehistoric meals, are enormous; some examples up to 0.5 km (600 yd) long, 100 m (300 ft) wide, and 8 m (25 ft) high are found along major rivers in the southeastern United States. Along the Atlantic seaboard, huge shell middens also accumulated, documenting the importance of marine foods in the diet of the hunter-gatherers of New England and the East Coast. In the Great Lakes region, native copper from the Lake Superior region was used extensively by peoples of the Old Copper culture. Copper knives, spearpoints, and various pendants and jewelry were cold-hammered from nuggets of native copper (Figure 5.3). Archaic sites in the Great Plains document a major focus on bison hunting. Sites in the dry, desert West contain artifacts and organic materials that indicate an emphasis on both plant foods and hunting for subsistence. Groups in the Great Basin collected numerous seeds and nuts and hunted antelope and small game such as rabbits. The location of the sites discussed in this chapter, as well as the timeline for these places, is shown in Figure 5.4.
A te tm m os pe ph ra er t u ic re
Page 167
e
7:59 AM
Ti m
10/1/09
Thousands of years ago
pri31057_ch05.pp164-197.qxd
Late glacial
PALEOINDIAN LATE PALEOLITHIC
Figure 5.2 The climate, geology, and archaeological sequence in Europe and North America during the past 12,000 years. The end of the Pleistocene is shaded.
Figure 5.3 Artifacts from the Old Copper Archaic culture in the midwestern United States. The longest spear point is approximately 15 cm (6 in) long.
We have been hunters for much of our past. Does that behavior have consequences today?
y _
Introduction The World after 8000 B.C.
167
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 168
Vedbaek Star Carr
Sannai Maruyama
Carrier Mills
PACIFIC OCEAN Equator
PACIFIC OCEAN
ATLANTIC OCEAN
INDIAN OCEAN Elands Bay Cave
Figure 5.4 Location of and timeline for selected settlements in the early Postglacial.
Star Carr Vedbaek Elands Bay Cave Sannai Maruyama Carrier Mills
10,000
8000
6000
4000 Years B.C.
_ _
168
Chapter 5 Postglacial Foragers
2000
0
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 169
Concept
<
The Postglacial Environment of Europe The conditions of the Present Interglacial
Holocene maximum
Temperature change (˚C)
2
Medieval climatic optimum
0 Little Ice Age
–2 Full Pleistocene
Younger Dryas
–4 16
12
8
4
0
Thousands of years ago
Figure 5.5 Average temperature increased dramatically from the late Pleistocene into the Holocene. The dotted line at 0°C represents the average temperature today.
0 Sea level (m below present)
Although the basic topography of the earth has not changed significantly since the end of the Pleistocene, the environment itself has undergone dramatic modifications in vegetation, fauna, and sea level. A marked shift in climate at the end of the Pleistocene was largely responsible for such environmental changes (Figure 5.5). Dramatic deviations from modern conditions are clear. Europe was as much as 8°C (20°F) colder during the ice ages some 18,000 years ago than it is today, and warmer than at present by 1–2°C (2–5°F) around 7000 years ago, during the early Postglacial. One of the results of the increasing temperatures at the close of the Pleistocene was the melting of continental ice sheets and a consequent rise in the level of the oceans. During the maximum cold period of the last glaciation, around 18,000 years ago, the sea was as much as 125 m (400 ft) below its present level. A gradual rise in the level of the oceans began after 16,000 years ago and continued during the early Postglacial (Figure 5.6). The rate of increase was variable, but a rise of as much as 1 m (3.3 ft) per century occurred during the period of maximum warming. The rising Postglacial seas did not reach present beaches until sometime after 5000 years ago. The higher sea levels of the Postglacial transformed the outline of the continents. Australia separated from New Guinea and Tasmania. The peninsula of Indonesia broke up into islands. The Bering Land Bridge was submerged and filled by the Bering Strait. The east coast of North America moved west by more than 100 km (60 mi) in some areas. The British Isles separated from the European continent as
20
40
60 10,000
5000
B.C./A.D. 1000
Year
Figure 5.6 Changes in sea level over the past 12,000 years. Holocene sea levels reached modern coastlines only in the past few thousand years.
_ Concept The Postglacial Environment of Europe
169
10/1/09
7:59 AM
Figure 5.7 Animals represented in the bones at the Mesolithic site of Star Carr in England, and the estimated weight of their meat. Notice that although the bones of roe deer are the second most common, the meat from this animal provides only a small part of the diet. Aurochs contribute the largest portion of animal protein.
Page 170
50 Number of identified specimens
45
Most common skeletal element
40 35 30 Percent
pri31057_ch05.pp164-197.qxd
25 20 15 10 5 0
Red Deer
Elk
Aurochs
the English Channel was flooded by rising seas. The forest history of Europe emphasizes the complexity of vegetational development through the end of the Pleistocene and during the early Holocene. Many of the plant species that disappeared during glacial conditions in the north survived along the southern coasts of Europe. Late Pleistocene deciduous forests were restricted primarily to southwestern Europe; hazel, oak, and elm survived in western France and northern Spain. Following the close of the Pleistocene, deciduous forests spread through most of western Europe, with the exception of small areas of Mediterranean forest in southern Italy, southern France, northeastern Spain, and the islands of the western Mediterranean. Along with those dramatic changes in climate and vegetation, there were pronounced changes in animal life. The large migratory herds of reindeer, mammoth, horse, and other game that roamed the tundra disappeared, either moving to the north where ice sheets previously existed or becoming extinct.
_ _
170
Chapter 5 Postglacial Foragers
Roe Deer
Wild Pig
In their place came the more recent species of European mammals adapted to the forest: the European elk (the same animal as the North American moose), the aurochs (wild cattle), the European red deer (the same animal as the North American elk), the wild boar, and the small roe deer. Those species made up the bulk of the terrestrial animal diet of Mesolithic hunters in Europe (Figure 5.7).
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 171
Site
:
Vedbaek Prehistoric communities in Mesolithic Denmark
Seabirds Ringed seal
Porpoise
Stranded animals
Algae
Herring Gray seal Cod
Plaice Flounder
Mollusks Crustaceans
Turbot
Porbeagle shark Mackerel
Denmark and the rest of Scandinavia have been occupied only briefly in the scale of prehistoric time, essentially since the close of the Pleistocene and the retreat of the ice sheets from northern Europe. At that time, as temperatures rose, the tundra gave way to open woodlands of birch and pine, and eventually to a mixed forest of lime, oak, elm, and other deciduous trees. These forests were occupied initially by herbivores such as aurochs and European elk, followed soon after by wild pig, the European red deer, roe deer, and many small mammals and birds. Inhabiting the streams and lakes were large numbers and varieties of fish. The inlets and islands of the seas around southern Scandinavia would have offered a rich source of food and were the locations of human settlement during the later Mesolithic. Wild animals and plants from the land, sea, and air were the focus of their hunting-andgathering activities. The inland forests were probably quite dense, supporting little wildlife. By 7000 years ago, there had been a dramatic shift in human social
Figure 5.8 A variety of marine foods were consumed by the inhabitants of Vedbaek. Some of those species and their preferred habitats are indicated here.
Figure 5.9 Excavations at Vaenget Nord. The excavators removed the topsoil, leaving the Mesolithic artifacts in place to map their location and distribution. Site Vedbaek
_
171
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 172
Figure 5.10 The plan of the Vedbaek site after excavation, showing specific activity areas and features. The site was located on a small island.
– 0.5
Bone dep deposits d posit siits
0.0
1.0
0.5 1.0
Lithicc refuse deposits p
1.5 2.0
2.5
Hide workin working
B Burial
Fli Flint working king
Primary ary occupation cu upa on Garbage midden en Cooking pit
N
– 0.5 5
Platform
0
10 m
Pits
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on an archaeological excavation in Norway, see the Internet exercises on your online learning center.
_ _
172
arrangements in this area from the first small, scattered groups of inland, tundra-dwelling reindeer hunters to concentrations of more sedentary societies along the coastlines. These groups expanded their resource base, eating fish, seals, porpoises, small whales, oysters, mussels, clams, and the like (Figure 5.8). Settlements became more permanent, and the dead began to be buried in cemeteries. An example of such a situation can be seen at an important archaeological area near the town of Vedbaek (vay-BEK) near Copenhagen, Denmark (Figures 5.9 and 5.10). Following the retreat of the ice, the Vedbaek Valley contained a freshwater system of lakes and streams. Warming trends continued, and rising sea level filled the mouth of the valley sometime around 5500 B.C., creating a brackish inlet. The shallow waters around the shoreline and islands of the inlet were covered with stands of reeds and sea grass. Over time, the inlet filled up with deposits of reeds, leaves, and other organic materials, becoming the layer of peat that it is today. In 1975, a Mesolithic graveyard was discovered here during the construction of a new school (Figure 5.11). The cemetery is radiocarbon-dated to approximately 4800 B.C. and contains the graves
Chapter 5 Postglacial Foragers
of at least 22 males and females of various ages. All the individuals in the burials were fully extended, with one slightly curled-up exception. Powdered red ochre was found in many of the graves. Racks of red-deer antler were placed with elderly individuals; males were buried with flint knives; females often were interred with jewelry made of shell and animal teeth. In one grave, a newborn infant was found buried on the wing of a swan next to his mother (Figure 5.12). The infant was buried with a flint knife, as were all the males in the cemetery. The mother’s head had been placed on a cushion of material such as an animal skin that was elaborately decorated with ornaments of snail shells and deer teeth. Similar materials were found around her waist, suggesting a skirt or costume of some kind. The cemetery also contained rather dramatic evidence for conflict among the groups occupying northern Europe at that time. The simultaneous burial of three individuals in a single grave—an adult male with a lethal bone point in his throat, an adult female, and a child—suggests both the violent death of all three and the existence of a nuclear family (Figure 5.13). Since the discovery of the cemetery, Vedbaek has been the focus of in-
10/1/09
7:59 AM
Page 173
N
pri31057_ch05.pp164-197.qxd
Woman and child Man
Old woman
Man Young adult
Old man Woman and infant Man Man, woman, Old man and child Empty grave Infant
Skull of young woman
Figure 5.11 The Mesolithic cemetery at Vedbaek, near the coastline of the former inlet. The darker area is the location of the construction activities that resulted in the discovery of the burials.
Woman
Young woman
Old man
Infant
Former coastline Saltwater inlet
tensive investigations. More than 50 archaeological sites have been located around the shore of the inlet, and excavations have been undertaken at several of these places. Some 60 species of fish, reptiles, birds, and mammals have been identified in the bone remains from these sites. These species come from every environment—the forest, streams, lakes, wetlands, the inlet, the sound, and the sea. Terrestrial animals are predominantly red deer, roe deer, and wild pig. Marine foods, however, provided a major portion of the diet; fish and seal bones are very common at the sites. Mesolithic sites in northern Europe were usually located on the shore, emphasizing the importance of the sea. Distinct zones of artifact deposition can be seen at such sites. The actual living floor on dry land is characterized by the presence of hearths, pits, construction stone, and some artifacts. Stone tools and other artifacts are generally small, suggesting that larger refuse may have been swept up and tossed or discarded elsewhere. Organic materials such as bone and plant remains generally do not survive on the surface of the ground in temperate climates. A second zone of refuse, originally discarded in the water next to the settlement, can be recognized in the layers adjacent to the occupation floor. Larger materials in this zone are well preserved, including stone, bone, antler,
and sometimes wooden artifacts. Because much of the shoreline of the Vedbaek inlet was occupied during the Mesolithic, this second zone often contains a vertical stratigraphy of tools and other debris. This information has been used to construct a detailed chronology for the area. Changes in artifact types and manufacturing methods can be traced through time. Repeated residence at the same location, however, tends to smear and obscure information about the horizontal arrangement of the prehistoric settlement, the locations of structures and associated hearths and pits. For that reason, an excavation was organized to uncover a settlement of brief occupation, where horizontal patterns
Figure 5.12 An artist’s interpretation of the mother and infant burial at Vedbaek. A photo of their grave opens this chapter.
_ Site Vedbaek
173
pri31057_ch05.pp164-197.qxd
10/1/09
Figure 5.13 The burial of an adult male, a small child, and an adult female from Vedbaek. Notice the lethal bone point in the throat of the male and the cluster of animal teeth on the chest of the female.
7:59 AM
Page 174
Bone Point
Tooth Pendants
of the use of living space might be examined. Several factors pointed to a site called Vaenget Nord (VING-it nord). Today, the location of this site is marked by a grove of birch trees growing on a slight rise in the landscape. The rise had been a small island during the period when an inlet of the sea filled the Vedbaek valley around 7500 years ago. The island was flooded and eventually submerged by rising sea level shortly after that date. Thus, the period when it could have been a platform for human occupation was limited. The age of the artifacts and radiocarbon dates reinforce these impressions. Excavations revealed that the number of artifacts per square meter was lower at this site than at the heavily used shoreline sites. Major excavations by teams of Danish and American archaeologists began in 1980 and concluded in 1983.
_ _
174
Chapter 5 Postglacial Foragers
The excavation strategy was twofold. Narrow trenches were cut into the deep marine deposits along the former shore of the island to reveal the refuse zone. Broad horizontal units were opened on the surface of the island to expose the living floor and the distribution of artifacts, pits, fireplaces, and other items. The surface sediments of the island are light sandy clays. The darker traces of past human activities such as digging, fire building, and the placement of posts are often retained in this light soil. On top of this natural surface of the island is a layer of cultural materials, made up of ash and charcoal, organic refuse, and the like. The thickness of this layer varies across the top of the island and is deeper along the sloping shoreline. In two areas at the southern and eastern margins of the island, large boulders had been fractured into numerous
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
pieces, creating a kind of pavement or landing area. In addition, there are several intact boulders on the site, concentrated in the northwest section. The excavations also exposed many very large posts. These posts are the trunks of elm or alder trees, roughly 30 cm (1 ft) in diameter, sharpened to a point and driven into the surface of the island. The distribution of flint artifacts on the island is of major importance to understanding the activities that took place at the site. The artifacts can initially be divided into the waste products of manufacturing and repair, and the actual finished tools that were in use. The finished tools are primarily adzes, arrowheads, and burins; only a few scrapers have been recovered. Waste materials include cores for producing flakes and blades, core-shaping flakes, and a great quantity of shattered flint as a by-product of the flaking technique. Most of the pits and hearths found in the excavation lie on the southwestern, landward, side of the island. Several large fireplaces are present, often associated with concentrations of burned cooking stones and charcoal. Also in this area is a large, shallow depression that may be the location of a structure of some sort. A dense cluster of small stakeholes was also observed in this area, perhaps related to the construction. Some small pits are scattered around the site. One human-size pit with a blade knife and two axes was probably a grave. The center of the island has several distinctive features, including the large depression, numerous pits, and stakeholes. Most of the projectile points come from this zone. This was likely the primary focus of the occupation. The northeast portion of the island contains evidence of hide working; most of the scrapers, truncated pieces, and unretouched blades come from this zone. Adjacent to this area is a zone of intensive stone toolmaking, rich in both débitage and flint tools. Most of the refitted pieces come from this area. To the south, behind these areas, is a zone of garbage dumping, containing abundant charcoal, fire-cracked stones, and only a few dispersed flakes. Along the north
Page 175
shore of the island is a zone of erosion, which was the beach during the time of occupation. North of the island, in what was the inlet, is the refuse zone of larger stone and organic materials discarded in the waters adjacent to the island. Density plots of the waste material reveal areas of tool manufacture. Flint axes were made as core tools by shaping a heavy nodule of flint into an elongated implement with a sharp leading edge. These axes, averaging 1 kg (2 lb) or more in weight, were attached to long elm handles and used to fell trees, to hollow logs, and to butcher meat. Arrowheads were made on rhomboidshaped segments of flint blades. Analysis of the microscopic wear on the edges of the stone artifacts indicates the function of certain tools and blades. About 25% of the blades show evidence for use on such materials as plants and wood, fresh and dry hide, and meat and bone. Distribution of the blades indicates the areas of the site where those activities took place. Hide-working tools were more common in central and northeast parts of the island. Vaenget Nord was probably a small and specialized camp for the Mesolithic inhabitants whose more permanent homes were along the shoreline of the inlet. This small island was likely used for certain activities at specific times of the year. Plant and animal remains indicate the utilization of a variety of environments, presumably during the warmer months of the year. Hazelnut shells were also common at the site, documenting fall habitation. The island situation and the presence of bones from garfish, mackerel, and dogfish indicate the importance of the inlet and the sea to the inhabitants. Vaenget Nord was the focus of activities that involved hunting (the presence of arrowheads), the butchering of animals (butchered bones, meat polish on blades), the manufacture of tools and equipment from animal by-products such as bone (wear marks on the burins) and hide (polish on scrapers and blades), and some woodworking (the presence of axes). The variety of hearths and pits and the heaps of cooking stones reinforce the impression of diverse activities.
_ Site Vedbaek
175
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 176
Concept
<
Bone Chemistry and Prehistoric Subsistence Information about past diet contained in human bones The nature of past human diet is one of the most important areas of prehistoric research. The quest for food directly affects many aspects of human behavior and society, including group size and social organization, residence patterns, technology, and transportation. Information on past diet has traditionally come from a number of lines of analysis: the study of preserved animal bones, plant remains, fecal matter, tooth wear and disease, and the physical characteristics of the human skele-
ton. New methods involving the chemical analysis of human bone provide a means of obtaining more information on paleonutrition. Human bone is composed of organic and mineral compounds and water. Isotopic studies of the composition of bone focus on the organic portion, primarily in the form of the protein collagen. The carbon atoms in collagen occur in two major stable forms, 12C and 13C (carbon-12 and carbon-13). The ratio of 13C to 12C (expressed as δ13C) in bone is
Wheat Average δ13C = –26.5 parts per thousand
Corn Human food based on 50% of each kind of plant
_ _
176
Chapter 5 Postglacial Foragers
Bone collagen δ13C = –21.4 parts per thousand
Average δ13C = –12.5 parts per thousand
Human food chain based on corn
Human food chain based on other plants only
Figure 5.14 The movement of carbon isotopes through the food chain. Differences in the types of plants consumed or the presence of marine foods in the diet will result in changes in the carbon isotope ratio (δ13C) in human bone. This information is used to estimate the diet of prehistoric human groups.
Photosynthesis in corn and some tropical plants
Atmospheric carbon dioxide δ13C = –7 parts per thousand
Photosynthesis in most plants
Bone collagen δ13C = –14.4 parts per thousand
Bone collagen δ13C = –7.4 parts per thousand
10/1/09
7:59 AM
Page 177
Neolithic
Marine animals
Mesolithic
Figure 5.15 Carbon isotope measurement of human bone from Mesolithic and Neolithic Denmark. A dramatic decline in the use of marine foods came with the onset of agriculture, around 4000 B.C.
–10
Terrestrial animals
–15
–20
δ13C
pri31057_ch05.pp164-197.qxd
–25
–30 6000
4000
2000
B.C./A.D.
Year
measured with an instrument known as a mass spectrometer and reported as a value ranging from about 0 to –30. The ratio of these two isotopes in bone reflects what is consumed in the diet. Carbon-13 is more common in certain kinds of terrestrial plants, such as corn, than in others and more common in the oceans than in other locations (Figure 5.14). People who eat corn have higher ratios of carbon isotopes in their bones. Changes in this isotope ratio in prehistoric bone can indicate when corn became an important component of the diet. Such studies have been done both in Mexico, to determine when corn was first domesticated, and in North America, to record when this important staple first arrived. Marine plants and the marine animals that consume those plants exhibit carbon isotope ratios ranging between –10 and –18 δ13C. Values more positive than –20 indicate a predominance of marine foods in human diet, whereas more negative values are more closely related to the consumption of terrestrial plants and animals. Henrik Tauber, of the National Museum of Denmark, measured the carbon isotope ratio in the bones of skeletons from Vedbaek, along with a number of other burials from Scandinavia and Greenland (Figure 5.15). The ratios from Vedbaek range from –13.4 to –15.3 δ13C and are close to values for
historical Eskimo skeletal material. Greenland Eskimo consumed marine foods extensively, making up perhaps as much as 90% of their diet, and a similar proportion of seafoods may have characterized the diet of the later Mesolithic. Saltwater fish, seals, porpoises, whales, and mollusks constituted a major part of the diet at Vedbaek. Measurement of a dog bone from the Vedbaek area gave a reading of –14.7 δ13C, within the human range, suggesting a very similar diet for the canine. Neolithic burials in Denmark show a sharp decline in carbon isotope ratios, indicative of a decrease in the importance of seafoods among the early agriculturalists. Similar patterns are noted elsewhere in Europe for the end of the Mesolithic. Nitrogen isotopes are used in much the same way as carbon isotopes but they provide different information about diet. The ratio of nitrogen-15 (0.37% of all nitrogen in nature) to nitrogen-14 (99.63% in nature) is used in paleodiet studies. Nitrogen is reported as δ15N, and values in human bone range from approximately +5‰ to +20‰. This nitrogen ratio is measured in bone collagen using a mass spectrometer. Variations in nitrogen isotope ratios are largely due to the role of leguminous plants in diet and the trophic level (position in the food chain) of the
leguminous plants Vegetables used
as food. trophic level An organism’s position
in the food chain.
Concept Bone Chemistry and Prehistoric Subsistence
_
177
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 178
organism. Atmospheric nitrogen (δ15N = 0‰) is isotopically lighter than plant tissues; values in soil tend to be even higher. Non-nitrogen-fixing plants, which derive all of their nitrogen from soil nitrates, can thus be expected to be isotopically heavier than nitrogenfixing plants, which derive some of their nitrogen directly from the atmosphere. These values in plants are passed through the food chain accompanied by an approximately 2–3‰ positive shift for each trophic level, including between mother and nursing infant. Grazing animals exhibit 15N enrichment, and more positive δ15N, com-
_ _
178
Chapter 5 Postglacial Foragers
pared to the plants they eat; predators show enrichment relative to their prey species. There are also differences in nitrogen isotope ratios between marine and terrestrial sources of food that can be used in the study of past diets. Human consumers of terrestrial plants and animals typically have δ15N values of 6–10‰, while consumers of freshwater or marine fish, seals, and sea lions usually have δ15N values of 15–20‰. Nitrogen isotope ratios may also vary with rainfall, altitude, and other factors. The nitrogen system, in general, is less well understood than that of carbon isotopes.
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 179
Site
:
Elands Bay Cave and De Hangen Late Stone Age hunter-gatherers in southern Africa
Much of what we know about the later prehistory of southern Africa has been discovered only in the past 30 or so years. During that period, John Parkington, of the University of Capetown, has been studying the artifacts and settlements of Postglacial groups in western South Africa. He says, We wish to know how the prehistoric populations used the area and its resources, what the relationship was between hunters and herders, what the exploitation strategies were, what precise resources were utilized and how, and what effects the resource pattern had on group size, site location, and seasonal mobility. Archaeology with an ecological viewpoint is the study of the interrelationships of technology, subsistence, and environment. (1984, p. 92) An important question Parkington and many other archaeologists face is whether prehistoric hunter-gatherers moved in an annual cycle of resource use and, if so, how far. Was their movement between the coast and inland in an east-west direction, or did they move north-south within environmental zones? Or were these groups relatively sedentary, moving only short distances when necessary? Those are difficult questions to answer. One problem is the relationship between the archaeological remains, primarily stone tools and animal bones, and the actual behavior of the prehistoric hunter-gatherers. A connection must be found between static, material objects and dynamic, human activities. Parkington has used several assumptions to make that connection. He says, Environments effectively comprise sets of resources, each with a specific distribution pattern. In order to extract and manipulate chosen resources, prehistoric groups needed to engage in specific activities, some of which involved the use of stone
Strandveld Fynbos
De Hangen
Karoo
Atlantic Ocean Elands Bay Cave
0
tools. Thus the accumulation of stone tools at a site reflects, albeit in a complex way, the range of activities pursued there. Further, differences between the composition of assemblages reflect the relative importance of tool-related activities at those sites. (1984, p. 96)
30 km
Figure 5.16 Environmental zones in southern Africa, and the site locations at Elands Bay Cave and De Hangen. The triangles indicate other known sites in the research area.
A review of Parkington’s investigations provides some indication of the nature of archaeological research. This project focused on an area of some 22,500 km2 (8700 mi2, about the size of the state of Vermont), between the western coast of the cape of southern Africa and interior mountain plateaus. Three major environmental zones running in north-south bands can be distinguished in the region (Figure 5.16). To the west, the Atlantic Ocean plays a major role in the climate and environment of the area. Many edible animals are available from the sea and the beach, including fish, seals, lobsters, mollusks, and seabirds. Rocky points near the mouths of rivers
Site Elands Bay Cave and De Hangen
_
179
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 180
Figure 5.17 Elands Bay Cave, on the west coast of South Africa.
transhumance A pattern of seasonal
_ _
movement usually associated with pastoralists who take their herds to the mountains in summer and to the valleys in winter; more generally, a regular pattern of seasonal movement by human groups.
180
and estuaries are particularly good locations for obtaining shellfish and other marine species. The Strandveld (STRAAN-feld) is a low, sandy band, up to 30 km (20 mi) inland, between the beach and the mountains. Vegetation in this zone produces fruit, berries, and roots in season. Animal species in the area include tortoise, dassie (similar to a hare), small antelope, and ostrich. Larger species of game such as the eland, a kind of large antelope, have disappeared since European contact. The middle Fynbos (FINN-boss) zone covers the sandstone mountains; shrubs predominate, but trees are found along rivers and at elevations over 1000 m (3000 ft). Plants and animals are abundant in this zone, and fish can be caught in the rivers. The Karoo (kaa-RUE) is a dry lowland zone between higher elevations, drained by the Doorn River. This area is a rich pasture for winter grazing by many species of herbivores. These same species may have moved west into the Fynbos during summer when the Karoo veld (grassland) was burned brown by the sun. Human hunters may have followed the herds in a similar pat-
Chapter 5 Postglacial Foragers
tern of regular movement, known as transhumance. Seasonal changes in resources are important for hunter-gatherers, who depend on wild plants and animals. Most of the edible fruits and berries in the research area ripen in late spring and summer and rarely appear during winter. Diaries and other records from early European explorers in southern Africa report that the diet of the hunter-gatherers in the inland was dominated by roots the women collected daily. Roots and tubers reach maximum size in late summer and are smaller during winter. Animal foods are generally available throughout the year, with some exceptions. Young seals are most easily captured along the coast during warmer months (August–November in the Southern Hemisphere) at breeding colonies on the beaches after their birth, or during winter when they come out of the water. Mollusks are better to eat during winter. Certain species of saltwater fish are available only seasonally on the coast. Water is also a concern in southern Africa. Precipitation is higher in the mountains. Smaller streams and freshwater springs are common in the mountains, but surface water is rare in the Strandveld and the Karoo. Several rivers run through the project area to the sea and provide some fresh water year-round. Parkington has suggested that the archaeological remains of huntergatherers found throughout this area were produced by the same or related groups of people moving seasonally in an annual cycle, from the coast to inland areas and back, to take advantage of resources when they were available. He argues that the Fynbos is best suited to summer residence, whereas both the Karoo and the coast provide better sources of food during winter. Shellfish and marine species substitute for the absence of plant foods during the cooler months of the year. The archaeological remains for evaluating that hypothesis come from excavations at two sites: Elands Bay on the coast and De Hangen (dee HAANGin) in the inland mountains. Elands Bay Cave is located at the mouth of the
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 181
Mole rat
Layer 1 2 3 4 5–7 8 9 11 12 10 13 15 16 17 18 19 20
C-14 Years B.P. 315 ±
50
± ± ± ± ± ±
85 80 45 85 80 90
1120 1520 3510 3780 7910 9600
Small bovid
Fur seal
Large medium bovid
Small medium bovid
Large bovid
Figure 5.18 Changes in the occurrence of animals by species over time at Elands Bay Cave. The layer numbers and radiocarbon dates are shown on the left side. Six species are shown: mole rat, fur seal, and four sizes of bovid (antelope). Of particular note is the increasing importance of fur seals and the trend toward smaller bovids in the upper layers.
10,700 ± 100 11,070 ± 140 12,450 ± 280
Vlie River at the Atlantic Ocean (Figure 5.17). Archaeological deposits from the early Postglacial consist primarily of large accumulations of shell, known as shell middens, which are filled with artifacts, bones, and plant remains. All of the evidence suggests that the inhabitants at Elands Bay were on the coast for a brief time during winter, exploiting a variety of marine and terrestrial animals (Figure 5.18). The shells of mussels and limpets are the most common food remains in the shell midden. Near the cave are a series of intertidal, rocky flats, where mollusks are still very abundant today. These shells suggest that this site was occupied repeatedly in winter, because the mussels can be toxic if eaten during summer. The shell midden also contains a variety of remains of both marine and terrestrial species. Marine fish, seals, and birds outnumber the terrestrial animals by ten to one. The presence of commercial lobster fisheries at the river mouth today confirms the abundance of these foods. Seal bones from Elands Bay Cave are similar in size
and came from yearling seals. These animals tire easily during their first period at sea and wash up or haul out commonly on the beaches during late winter. Dassie bones are also present at Elands Bay Cave, but not in abundance; most of the individuals are fully adult, and newborns are rare. Some of the dassies are juveniles, with newly erupting permanent teeth. In modern animals, this dental growth occurs in September and October, the end of the winter season in the Southern Hemisphere. Although grasses used for bedding, as well as seaweed and twigs, are all preserved in the shell midden, there is very little evidence for the use of plant foods. De Hangen is located in the Fynbos, some 80 km (50 mi) from the coast. Excavations at the site have also produced a variety of plant and animal remains. Long grasses were gathered during late spring and summer for bedding in the caves. Edible plant foods were very common, including seeds, fruits, tubers, and roots, more readily available in the summer months.
shell midden A mound of shells ac-
cumulated from human collection, consumption, and disposal; a dump of shells from oysters, clams, mussels, or other species found along coasts and rivers, usually dating to the Holocene. adze A heavy, chisel-like tool.
Site Elands Bay Cave and De Hangen
_
181
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 182
The animal bones came primarily from tortoises, freshwater fish, dassies, and small antelopes. Most of the dassies were adult or very young when they were killed, during summer. Connections with the coast are also in evidence. Several marine shells were found. Rock paintings in the mountains south of De Hangen depict sailing ships, which must have been observed from the coast (Figure 5.19). Stone tools were made from a material available only from sources in the Strandveld. Such tools provide additional information about the pattern of residence. There are major differences in the lithic assemblages of the coast and the inland mountain area. Adzes, associated with woodworking, constitute more than half the tools at sites in the mountains and only about 5% of the assemblages on the coast. Adzes were probably used to make sticks for digging roots and tubers in the Fynbos area. Pieces with one blunted edge are rare in the mountains, more common in the Strandveld. These small stone artifacts were likely used as sharp tips and barbs for arrows, and thus were found more often where hunting is the primary means of subsistence.
Figure 5.19 A cave painting of a sailing ship from one of the inland sites in the De Hangen research area, evidence of contact with the coast.
_ _
182
Chapter 5 Postglacial Foragers
The archaeological remains suggest a general pattern of winter residence on the coast and summer residence in the mountains, supporting Parkington’s hypothesis. This evidence also fits the pattern that might be expected given the seasonal availability of certain resources. However, the archaeological evidence does not prove this to be the case. Settlements could as easily have been moved a few tens of meters as a few tens of kilometers. Winter sites inland, or summer sites along the coast, simply may not yet have been discovered. Bone chemistry provides contradictory information (see “Bone Chemistry and Prehistoric Subsistence,” p. 176). If the same individuals were moving seasonally from inland to coast, eating seafood for at least part of the year, a marine diet should appear in the carbon isotope content of their bones. However, analysis of human skeletal material from inland burials has failed to reveal the expected isotope ratios, suggesting that visits to the coast were not part of a regular pattern.
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 183
Site
:
Sannai Maruyama Hunter-gatherers of Jomon Japan
Archaeology is a national craze in Japan, where more than $1 billion a year is spent on excavations at more than 13,000 sites. (In contrast, the amount spent for archaeology by the National Science Foundation in the United States is around $9 million annually.) The early Postglacial period in Japan is a long and well-documented sequence of environmental change and cultural development. Dating to 10,000–300 B.C., this period is referred to as Jomon (JOEmoan). Over 35,000 Jomon sites have been discovered, concentrated in the eastern part of the country. Subsistence was based primarily on wild resources, although some cultivated plants may have been included in the diet. The importance of hunting is indicated by the
abundance of arrowheads at many Jomon sites (Figure 5.20). Inland hunting focused largely on deer and boars, but bears, antelopes, monkeys, hares, martens, and other small mammals were also taken. Marine foods and coastal settlements predominated in eastern Japan, particularly after 4000 B.C. Fishing was important at sites on the rivers, on estuaries, and on the ocean. Elaborate fishing gear, and even whole communities specializing in fishing, is evidenced in the archaeological record. Salmon and other freshwater fish were caught in the rivers. Bones from such deepwater species as tuna and shark were also found in the deposits. Fishing equipment included net floats of
Hunting
Gathering
SPRING
For a Web-based activity on the history of ancient Japan, see the Internet exercises on your online learning center.
Figure 5.20 The Jomon annual cycle of hunting and gathering. The chart depicts the relative importance of different food resources and subsistence activities during the seasons of the year, with domestic activities in the center. The widths of the bands indicate the importance of the various activities.
WINTER
Seafood collecting
w w w. m h h e . c o m / p r i c e i p 6 e
House building Stone toolmaking
AUTUMN
Pottery making
Fishing and sea mammal hunting
SUMMER
Jomon The archaeological culture of late-Pleistocene and early-Holocene Japan.
Site Sannai Maruyama
183
_
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 184
C
C
B A
Pot Buried pot Milling stone Mortar
D
Polished stone axe A
D
B
Flaked stone axe Stone projectile point
0
Figure 5.21 The plan and sections of a Middle Jomon circular pithouse from central Honshu, Japan.
y
Can prehistoric groups be both hunter-gatherers and farmers?
cord-marking A decorative technique
in Jomon Japan and elsewhere, in which cord or string is wrapped around a paddle and pressed against an unfired clay vessel, leaving the twisted mark of the cord.
_ _
pithouse A prehistoric semisubterranean dwelling in which the lower parts of the walls were the earthen sides of a shallow pit; the top parts of the walls often consisted of a framework of poles intertwined with small twigs, covered with mud.
184
3m
bark and pumice, net weights, fishhooks, harpoons, and dugout canoes. Over 30 species of shellfish were found in the shell middens. Plant foods were also important in the diet, and more than 180 species have been identified from Jomon settlements, including fleshy fruits such as grapes, blackberries, and elderberries; aquatic plants with edible seeds; and tubers. Nuts played a major role in subsistence; the shells of walnuts, chestnuts, acorns, and buckeyes occur at many sites. The preparation of nuts, seeds, and roots is indicated by the presence of mortars, pestles, milling stones, and other equipment. In the later Jomon, root crops such as taro, yams, and arrowroot may have been under cultivation in western Japan, along with some cereals, such as buckwheat and barley. Chipped stone hoes were common at sites from this time. Rice, however, was apparently not cultivated in significant quantities until toward the end of the Jomon period. Wood, bone, and antler were used for various tools and other objects; these items were often heavily decorated with carved designs. Jade was exchanged over long distances to be
Chapter 5 Postglacial Foragers
manufactured into pendants and other forms of jewelry. Pottery containers were invented very early in Japan, probably before the end of the Pleistocene. These fired clay vessels occur often at Jomon sites, and their forms became increasingly elaborate over time. Distinct regional varieties can be identified. The word jomon means cord-marking, referring to the kind of decoration applied to pottery before firing. Cord-wrapped sticks were pressed into the wet clay to create a design. Jomon villages varied greatly in size and often were occupied for a long time—several hundred years or more, in some cases. Settlements along the coastline are associated with shell middens. Inland sites are found on higher, level ground. The inland settlements generally contained 4–8 pithouses, although up to 50 structures from a single site have been identified. Pithouses were substantial circular or rectangular structures built of large support posts, containing elaborate, stone-lined hearths (Figure 5.21). Villages were often arranged in a horseshoe shape and may have been divided into two groups of residential units, one on either side of the horseshoe. Communal structures for various purposes were built around the periphery of the village. Burials occurred individually near houses or in the communal structures. The central open area contained few artifacts, but standing stones were occasionally erected there. In one instance, these large standing stones, 1.5 m (5 ft) high, were transported from a quarry some 120 km (80 mi) away. The site of Nittano (NI-ta-no), near the modern city of Tokyo, is fairly typical of Jomon fishing communities along the coast. The site is marked by an extensive shell midden and was occupied during the Middle Jomon period, about 4000–3000 B.C. Some 23 species of mollusks and 13 species of fish were identified. Analysis of mollusk shells from Nittano indicates that the site was occupied year-round. The shells of mollusks such as oysters and limpets grow in a series of increments, represented in the shell as daily growth rings of calcium carbonate. Using a
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
microscope, researchers can count and compare these rings to the known month of birth for the species, thereby determining both the age of the mollusk and the time of year it was collected. Most of the mollusks at Nittano were harvested during spring and autumn, a few during summer, and almost none during winter. Other sites document the diversity of Jomon adaptations and the possibility of farming activities in this period. Located in Aomori Province in northern Japan, the site of Sannai Maruyama provides an extraordinary picture of life during the Jomon period. The site is located inland, about 20 m (60 ft) above sea level, along the Okidati River. Excavations have been conducted at the site off and on since 1953 (Figure 5.22). The site covers more than 60,000 m2 in an area roughly 250 m in diameter, about eight city blocks. Approximately 40% of the site has been excavated. The Jomon settlement here lasted from 5900 to 4300 B.C., almost 1600 years, and must have held hundreds of inhabitants. The archaeology of the site is remarkable in many ways. The evidence takes the form of architectural remains, features, and artifacts. The varied remains include more than 500 pit dwellings of two types. Many small, residential pithouses are semisubterranean and circular, 3–4 m in diameter, and probably were family residences. There are also a few very large, long pithouses, one up to 32 m in length. These may have functioned as meeting places or workshops. Enormous storage pits (Figure 5.23) were also found in the excavations. A number of large towers or shrines were built using six huge posts as a foundation. The posts of one of these structures were chestnut tree trunks up to 1 m in diameter. In addition, there are stone circles associated with adult burials. Adult graves are also found along the three major roads that cross the site. These roads are several hundred meters in length and up to 40 m wide. Children were buried in ceramic jars around the pithouses. Much of the artifactual evidence from the site came from two large refuse
Page 185
mounds and waterlogged deposits in a wetland at the edge of the settlement. Artifacts were commonly either ceramic sherds or stone tools. The elaborate, cord-marked pottery was found in a range of shapes, used as cooking vessels, plates, and bowls (Figure 5.24). Clay for the ceramics was quarried from deposits at the edge of the site. The stone tools included arrowheads, axes, and grinding stones. Some of these stone artifacts were made of obsidian imported from almost 600 km to the south. Jade was also imported for beads and jewelry, coming from a source 500 km to the north. Amber and bitumen (a natural asphalt adhesive) were transported to the site from 200 km distant. Among the many remarkable finds in the waterlogged deposits were dugout canoes, baskets, rope, and textile
Figure 5.22 A trench through the south mound at Sannai Maruyama showing the stratigraphy.
_ Site Sannai Maruyama
185
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 186
Figure 5.23 An excavated storage pit at Sannai Maruyama, with a depth of almost 2 m (6 ft).
Figure 5.24 Pottery vessels from the site of Sannai Maruyama.
fragments made from plant fibers. In addition, several pieces of lacquerware, along with seeds from the varnish tree and the tools used to produce these spectacular objects, were found. Lacquerware is made by heating the sap of the varnish tree and applying many coats of the lacquer to a wooden bowl or jar. Lacquerware is resistant to moisture and heat and has a durable color and luster, adding beauty to the object. This is a complex process and suggests that craft specialists were present at Sannai Marumaya. The plant and animal remains at the site document diet and nonfood re-
_ _
186
Chapter 5 Postglacial Foragers
sources. Huge quantities of chestnut shells reflect the importance of this tree for food, building material, and firewood. Pollen evidence indicates that chestnut quickly replaced other species in the neighborhood of the settlement, suggesting that the inhabitants were actively cultivating these trees. There is some evidence for cultivation during the Jomon in some parts of Japan, involving domesticated gourds (not a native species) and beans, and perhaps millet, beginning in the Early Jomon. Phytolith evidence has been used to suggest rice cultivation in western Japan in this same period. The faunal remains contained a combination of terrestrial and aquatic species. Rabbits and flying squirrels were the most common land mammals; ducks and geese dominated the bird bones. The sea was an important source of food, and the bones of fish (especially yellowtail and shark) and marine mammals such as small whales and sea lions were present. The people of Sannai Maruyama also ate blowfish, a highly poisonous species that must be processed very carefully to remove the toxins. Today Sannai Maruyama is a protected historical park (Figure 5.24). Various aspects of the archaeology of the site have been preserved in situ or in a museum at the site to show the achievements of the Jomon people to schoolchildren, local people, and tourists from all over Japan.
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 187
Site
:
Carrier Mills A Middle Archaic settlement in southern Illinois
In 1979, the Archaeological Resources Protection Act provided for the protection of archaeological resources located on public lands and Indian lands; defined archaeological resources to be any material remains of past human life or activities that are of archaeological interest and are at least 100 years old; encouraged cooperation between groups and individuals in possession of archaeological resources from public or Indian lands with special permit and disposition rules for the protection of archaeological resources on Indian lands in light of the American Indian Religious Freedom Act; provided that information regarding the nature and location of archaeological resources may remain confidential; and established civil and criminal penalties, including forfeiture of vehicles, fines of up to $100,000, and imprisonment of up to 5 years for second violations for the unauthorized appropriation, alteration, exchange, or other handling of archaeological resources with rewards for furnishing information about such unauthorized acts. (U.S. Congress, Office of Technology Assessment, 1986) Before the start of federally funded construction, the U.S. Environmental Protection Agency requires an environmental impact study to determine whether important archaeological or historical sites are in danger of destruction. Both federal agencies and private corporations must provide information on the effect their projects may have on the history and prehistory of the area. Various kinds of construction—reservoirs, highways, sewage systems, power lines, to name just a few—require such archaeological surveys and environmental impact statements. This kind of work, known as cultural resource management, is an important part of the field of archaeology. Strip mining completely removes the surface of the ground in order to
excavate deep pits to extract the coal. Any buried or surface archaeological sites that are not rescued in some way before strip mining begins will simply be destroyed. In the early 1980s, the Peabody Coal Company of St. Louis, Missouri, was required to submit an environmental impact study of the Carrier Mills area in southern Illinois before expanding strip-mining operations there. The area to be surveyed and tested for archaeological and historical sites covered some 57 ha (150 acres). Richard Jeffries, Brian Butler, and David Braun, of the Center for Archaeological Investigations at Southern Illinois University–Carbondale, were the directors of this 5-year project. The survey of the Carrier Mills area discovered several large multiperiod sites, dating to between 4000 and 1500 B.C. (Figure 5.25). These sites were excavated carefully to obtain as much information as possible before the mining began. One of the three settlements, known as the Black Earth site,
Figure 5.25 The location of the Black Earth site, Sa-87, at Carrier Mills, Illinois. The dark areas on the ground correspond to the settlement areas at the site.
cultural resource management The
survey and/or excavation of archaeological and historical remains threatened by construction and development.
_ Site Carrier Mills
187
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Figure 5.26 The locations of excavation units in the three midden areas of the Black Earth site.
Page 188
Midden stains
N
0
50 m
Area B
Area A Area C
Extraction tasks that have been identified for the Middle Archaic component include hunting (projectile points, atlatl components, faunal remains), fishing (fishhooks, fishhook production residue, faunal remains), fowling (faunal remains), procurement of lumber (axes), nut procurement and processing (pitted stones, grinding stones, nutshell), butchering (heavy-duty scrapers, cutmarks on bone), and hide processing (scrapers, bone awls). Maintenance tasks include the manufacture of flaked stone tools (unfinished bifaces, débitage, hammerstones), ground stone tools (unfinished tools), and bone and antler tools (grooved abraders, antlers with scoring marks, production residue), and the recycling of broken tools (scrapers made from broken projectile points). One socially related task can be accounted for in the data, disposal of the dead (burials). —Richard Jeffries and Mark Lynch (1985)
_ _
midden An accumulated pile of trash and waste materials near a dwelling or in other areas of an archaeological site.
188
dates to the Middle Archaic, 4000– 3000 B.C. The site sits along a low ridge, next to a series of shallow lakes and swamps to the east and west; the low, rolling Shawnee Hills lie to the south of the site. The occupants used the river, lakes, and swamps of the lowlands for aquatic resources and exploited the uplands behind the settlement for deer. The name for the site comes from the distinct black middens—accumulations of trash, waste materials, and charcoal and ash from thousands of campfires—that stand out against the yellow-brown soils of the ridge (Figure 5.26). The entire scatter of artifacts at the Black Earth site covers some 53,000 sq m (13 acres). It is an enormous site, more than a city block in size. Three discrete concentrations were recognized within the larger area, designated A, B, and C. The largest, Area A, is at least 17,000 sq m (4 acres) (the size of two soccer fields), and the cultural horizon extends to 1.5 m (5 ft) below the present ground surface. Evidence for the use of these areas and for the activities of the inhabitants comes from excavation of the prehistoric middens at the site (Figure 5.27). Deposits in the midden areas appear to be continuous, suggesting that the occupation was relatively uninterrupted. The middens contain much more sand than the surrounding natural soils, suggesting that locally available sand-
Chapter 5 Postglacial Foragers
stone may have been brought to the site for various uses. The decomposition of this rock would explain the higher levels of sand in the midden soils. The concentration of ash, excrement, and other organic matter in the sediments changed the chemistry of the soil, resulting in more alkaline conditions, favorable for the preservation of bone and other materials. Preserved plant and animal remains document the diet of the inhabitants (Figure 5.28). Almost 57,000 pieces of animal bone were excavated, representing 77 species of mammals, reptiles, birds, and fish. Turtles from the nearby swamp and wetlands are among the most common remains. The bowfin, favoring swampy habitats, is the most common fish represented. White-tailed deer and turkey are the most important land animals. Raccoon, opossum, elk, rabbit, beaver, and other species were also included in the game larder, but deer must have been the staple. Deer were hunted throughout the year, but more animals were killed during autumn and winter. Nonfood products, such as fur, feathers, and shells, were also important. Hickory nutshells are very common in the middens; acorn shells also occur throughout the occupation area, along with the remains of hazelnuts and walnuts. These nuts are available in autumn and could easily be stored and used throughout the winter for
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 189
food. Seeds from hackberry, wild grape, persimmon, bedstraw, and wild bean are also common. These plants ripen and are harvested during late summer and fall. Discarded stone artifacts in the middens include a number of hafted tools: projectile points, drills, and end scrapers. Scraping tools were often made from broken arrowheads, recycled for cleaning animal hides. Ground stone tools are also common, ranging from grooved axes to grinding slabs, and from counterweights to spearthrowers to hammerstones. Bone and antler tools are numerous and include awls, pins, needles, fishhooks, and other items. Almost one-third of these were made from the antler and bone of white-tailed deer. Turtle shells were used for bowls, cups, and dance rattles. Several recovered bone pins may have been used with clothing or in hair
TABLE 5.1 Grave Goods Buried with Individuals at the Black Earth Site Function
Adult Male
Adult Female
Child
Utilitarian
Mussel shell Bone awl Worked antler Unworked deer bone Projectile point Chert core Drill Scraper Utilized flake Axe fragment Shell pendant Bone pin Anculosa-shell bead Antler bead Plummet Bar gorget Decorated bone Animal bone (not deer) Elk antler cup Red ochre Fossil Fluorspar crystal Worked shale Worked stone Hematite Water-worn pebble Banded slate Miniature grooved axe
Mussel shell Bone awl Worked antler Unworked deer bone Bone needle Hafted scraper Chert biface
Mussel shell Utilized flake Ground stone
Shell pendant
Shell pendant Shell bead Mussel-shell bead Circular shell disk Crinoid-stem bead
Worked turtle shell Animal bone (bird) Red ochre
Turtle shell cup Red ochre Water-worn pebble Fluorspar crystal
Ornamental
Ceremonial
Figure 5.27 Excavations at the Black Earth site in the 1970s exposing pits and graves.
_ Site Carrier Mills
189
pri31057_ch05.pp164-197.qxd
10/15/09
10:51 AM
Page 190
Figure 5.28 An artist’s interpretation of Middle Archaic life at the Black Earth site.
y
What are the consequences of sedentism for human society?
grave goods The items that are
placed in graves to accompany the deceased.
arrangements. These pins are decorated with incised lines in a geometric design. Beads and pendants of shell, antler, and bone also were found. In addition to the settlement, there was a cemetery at the Black Earth site. At least 154 burials from the Middle Archaic were found. The investigators estimate that there were more than 500 graves in the entire site. Infants accounted for 21% of the burials. Males slightly outnumbered females in the cemetery and had a shorter lifespan. Male life expectancy after childhood was only 32 years; for females, 38 years (Figure 5.29). Adult males averaged 1.7 m (5 ft, 6 in) in height; females were 1.6 m (5 ft, 2 in). Individuals were buried in shallow pits (Figure 5.30), and occasionally a cap of clay was used to seal the grave. About half the burials were found in an extended position, and the other half were curled up. Children and young adults were usually fully extended in the graves, but older adults were placed in the graves in a
_ _
190
Chapter 5 Postglacial Foragers
flexed, or curled-up, position. Four individuals were buried in unusual positions, including one man seated upright and bent over. Grave goods provide an important sample of the tools and ornamental items of the living population (Table 5.1). Everyday objects were generally discarded when they were broken or no longer useful. Materials placed in graves, however, were likely the property of the deceased or placed with the dead as a kind of offering; these objects are intact when discovered at a burial site (see “Grave Offerings,” Chapter 7, p. 293). Only about 25% of the individuals were buried with objects in their graves. Males were buried with a wider array of goods than females or children. Males were generally buried with equipment to obtain food or other resources, and females were buried with tools to process or prepare those raw materials. Women were buried with scrapers, bone awls, and needles. Only males were buried with ornamental objects such as bone pins and decorated bones, along with projectile points, worked stone drills and axes, and small animal bones. A “medicine bundle” was placed near the head of a 43year-old male. This cluster contained 45 objects, including eagle talons, part of a bear paw, a miniature axe, pieces of slate, red ochre, and a dog’s tooth. One perforated disk of marine shell, probably from the Atlantic Ocean or the Gulf of Mexico, was found around the neck of a buried infant. Copper from the Great Lakes area was found with an adult male at Carrier Mills. A copper wedge had been placed at the top of the neck of the skeleton, perhaps as a substitute for the missing skull of this particular individual.
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 191
Age in years
Females
Figure 5.29 The demographic distribution of the Black Earth burial population by age and sex.
Males
55–60 50–55 45–50 40–45 35–40 30–35 25–30 20–25 15–20 10
5
0 Percentage
5
10
Figure 5.30 Burial 102 at the Black Earth site, with a cache of artifacts near the right shoulder.
_ Site Carrier Mills
191
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 192
Concept
<
The Human Skeleton A storehouse of information Cranial bones
Figure 5.31 The major bones of the human skeleton.
Facial bones Mandible Clavicle
Scapula
Sternum
Humerus
Ribs
Ulna Radius
Vertebrae
Carpals
Innominates
Metacarpals
Femur
Sacrum
Phalanges Patella Tibia Fibula Tarsals Metatarsals Phalanges
TABLE 5.2 Estimating Stature from Femur Length Stature
_ _
Femur Length, mm
cm
in
452 456 460 464 468
169 170 171 172 173
66 66 67 67 68
Source: Adapted from R. Jurmain, H. Nelson, and W. A. Turnbaugh. 1987. Understanding physical anthropology and archaeology, 4th ed. St. Paul, MN: West.
192
Since the time of the Neanderthals, humans have buried their dead. As a result, intact human skeletons (Figure 5.31) are sometimes preserved in single graves or cemeteries. Archaeologists and physical anthropologists spend hours carefully uncovering and removing these skeletal remains from the ground. Bones and teeth are taken to the laboratory, where a surprising amount of information can be obtained from the observation and measurement of various features. Consideration of a series of burials from a single cemetery can provide information about the entire population, including demographic
Chapter 5 Postglacial Foragers
characteristics, life expectancy, and inheritance patterns (Figure 5.32). The human skeleton is much more than a structural framework for supporting the body. Bone tissue contains a wide variety of information about the individual to whom it belongs. The length and thickness of bone, for example, provide an indication of an individual’s size and strength. Stature can be estimated from measurements of the long bones of the arm or the leg (Table 5.2). Evidence of disease or illness is also embedded in bone. It is often possible to determine age at death, the cause of death, sex, a history of disease or accident, and nutritional status from analyzing prehistoric human bone. Diet, too, may be reflected in the chemical composition of bone (see “Bone Chemistry and Prehistoric Subsistence,” p. 176). The sex of a skeleton is indicated by several features. Adult males are generally taller than females, but in most cases, size alone is not a sufficient indicator of sex. Several features of the pelvis are most important for determining the sex of skeletal remains. The female pelvis is wider than that of the male and slightly different in shape. Several indicators are used to estimate age at death, one of the most important being teeth. The age of children and adolescents is relatively simple to determine from teeth. In many mammals, the deciduous teeth of young individuals are replaced by permanent teeth, and the eruption patterns of permanent teeth are well established. For example, most 6-year-old humans are missing their two front teeth; we speak of a 12-year molar, referring to the age at which that tooth erupts. The size and condition of skeletal bones can also be used to estimate age at death. Body size, as indicated by the length of long bones, is one clue. How-
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 193
Figure 5.32 The skeleton as a storehouse of information. (1) Lesions on the skull suggest scalping at the time of death. (2) Flattened ribs and a fused breastbone result from corsets of the eighteenth century. (3) Badly twisted and eroded joints are evidence of arthritis. (4) Bone spurs appear on the knees of riders who spent too much time in the saddle. (5) Bone cells, or osteons, are denser in middle age (right) than in adolescence (left). (6) Shortened and deformed foot bones document the practice of foot binding in the China of previous centuries. (7) Osteoporosis, a loss of bone density, is more common in the elderly (bottom) than in the young (top). (8) The female pelvis is wider than that of the male, for giving birth. (9) More robust femurs are evidence of a diet higher in meat; thin bones suggest a proteindeficient diet. (10) Bumps and irregularities in bone show improperly healed fractures and injuries. (11) Indentations and grooves in the teeth are found among fishermen and wool spinners, who used their teeth for cutting thread.
11 1
10
2
9 3
8 4 7 5 6
ever, the joint ends of long bones, called epiphyses (e-PIFF-a-SEZE), provide even more reliable information. During childhood and adolescence, as the skeleton grows and hardens, the ends of long bones are separate from the shaft to facilitate growth. However, at puberty and during adolescence, the epiphyses harden and unite with the shaft to become a single bone, terminating growth. The age of epiphysis union varies for different bones, and that information can be used to determine the age at death. By the early twenties, the development of teeth and bones is largely complete in humans. Less precise indicators must be used to estimate the age at death from the skeletal material of older individuals. Skeletal features used to determine the age of older adults include the pelvis and the skull,
and degenerative changes in bone tissue. Stress during one’s lifetime is also revealed in the skeleton. Malnourishment in childhood causes the disruption of bone growth, which shows up in the skeleton as a series of distinct features, known as Harris lines, in the ends of long bones. Tooth enamel also reflects childhood stress and malnourishment in an irregular series of lines, a condition known as dental enamel hypoplasia (HI-po-PLAY-sha). Various diseases and other health problems also are reflected in bone. Arthritis is one of the most common conditions in the bones of older adults, characterized by an accumulation of bone tissue around the joints. Syphilis, tuberculosis, and other infectious diseases may result in bone loss and pitting or deformation of the skull and other bone surfaces.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on a Mesolithic burial in Italy, see the Internet exercises on your online learning center.
epiphysis The end of a long bone in humans and other mammals that hardens and attaches to the shaft of the bone with age.
Concept The Human Skeleton
193
_
pri31057_ch05.pp164-197.qxd
10/1/09
7:59 AM
Page 194
Concept
<
Contemporary Hunter-Gatherers A lost way of life at the margin of civilization
y
Are there any environments where hunter-gatherers didn’t live?
Figure 5.33 The seasonal diet of the Copper Eskimo.
About 99.9% of our human past was spent as hunter-gatherers—living in small, simple, egalitarian societies, moving often following a seasonal cycle in the quest for wild foods. For example, the Copper Eskimo of northern Canada followed a seasonal round of residential moves that focused on seal hunting in winter, fishing in summer, and caribou hunting in autumn (Figure 5.33). Some kind of an annual cycle defined the basic pattern of subsistence and settlement for many groups of hunter-gatherers. Group size and organization were highly flexible among hunter-gatherers and permitted a rapid response to seasonal changes in the food supply. Such adaptations are thought to represent the foundations of human society, from which more recent social and political arrangements have evolved. Among these groups, leadership was ephemeral and gained through the recognition of individual abilities. Relationships between individuals were determined by kinship; status differences were negligible; everyone had access to the resources of the group. Private property was limited; labor, goods, and food were shared communally. Essential tasks for food collecting were divided according to sex; men were hunters of larger animals, while women collected plant foods and fish and trapped small animals. Such groups survived until very recently in the more remote parts of the
world. Today, due to the spread of farming societies and the reach of the industrialized world, hunter-gatherer societies have basically been eliminated. The last of these groups lived in the most marginal environments— only those areas that were of little interest or use to agricultural and industrial societies. The San, or Bushmen, of the Kalahari Desert of South Africa (Figure 5.34); the Mbuti (MIMbo-tee) Pygmy of the Congo rain forest in Africa; the Hadza (HOD-zah) in the tsetse fly–ridden areas of East Africa; the Inuit, or Eskimo, of North America and Greenland; and the Aborigines of western Australia were among the few remaining groups of hunter-gatherers. Today, even those societies have abandoned their traditional way of life, since national governments have settled them in permanent communities for purposes of education and health care. The end of the twentieth century marked the demise of the hunting-andgathering way of life (Figure 5.35). Our picture of hunter-gatherer lifeways is too simple, however, and reflects relatively recent changes in human society. Ten thousand years ago, hunter-gatherers occupied all the continents except Antarctica; everyone was a hunter-gatherer. Theirs was a very successful adaptation, one that had provided for the enormous expansion of the human species into a wide range of environments all over the world.
Seal Polar bear Fish Caribou
_ _
Jan.
194
Chapter 5 Postglacial Foragers
Feb.
Mar.
Apr.
May
June
July
Aug.
Sep.
Oct.
Nov.
Dec.
pri31057_ch05.pp164-197.qxd
10/1/09
8:00 AM
Page 195
10,000 B.P.: World population 10 million, 100% hunter-gatherers
Figure 5.34 Kua hunter-gatherers of the Kalahari Desert of Botswana, South Africa. This woman is harvesting nutritious wild mongongo nuts.
A.D.
1500: World population 350 million, 1% hunter-gatherers
Inuit
Pygmy
Hadza
San
Aborigines
A.D.
1985: World population 4 billion, 0.0001% hunter-gatherers
A.D.
2000: World population 6 billion, 0% hunter-gatherers
Figure 5.35 The estimated population and distribution of hunter-gatherers over the past 10,000 years.
_ Concept Contemporary Hunter-Gatherers
195
pri31057_ch05.pp164-197.qxd
< ( : < V : y < p : W
10/1/09
The World of Hunter-Gatherers
H
unter-gatherers today occupy only the most marginal places on earth and have essentially disappeared. Hunter-gatherer societies exhibited a diverse range of forms, from large, sedentary populations with hierarchical leadership to very small, egalitarian groups moving about the landscape in search of food. In most areas, wild food was readily available to support large numbers of people. Such adaptations have been described as the “original affluent society” by anthropologist Marshall Sahlins (1968, p. 89). Leisure time was plentiful, and life must have been reasonably good. Health conditions were significantly better than among early farming populations. By the early Postglacial, many of these foraging groups had evolved into larger groups, with more sedentary occupations, more complex technology, and more intensive subsistence activities. Developments in technology, settlements, and subsistence practices are preserved in the archaeological record. Technology promoted greater efficiency in transport, in tools, and in food procurement. Watercraft and paddles are found at sites in a variety of coastal and lakeside settlements. The inhabitants of Franchthi (FRAUNK-tee) Cave in mainland Greece were sailing to the island of Melos (ME-los) by the end of the Pleistocene. Mesolithic coastal foragers in southern Scandinavia depended primarily on the sea for food. Settlements were generally larger, of longer duration, and more differentiated than in the preceding Paleolithic. Sedentary occupations are known from a number of areas, including Jomon Japan, the Middle Archaic in North America, and the Mesolithic of northern Europe. Domestic dogs were used for hunting in Mesolithic Europe and in North and South America. Food procurement was both more specialized and more diversified—specialized in the technology and organization of foraging activities, and diversified in the numbers and kinds of species and habitats that were exploited. Plant foods became more important for subsistence, and nutshells are especially visible in the archaeological record from this period. Fishing and the exploitation of coastal and aquatic resources became very important in the quest for food. Clearly, increasing sophistication in subsistence and settlement were important trends among Postglacial foragers. In certain areas, the changes among hunter-gatherers at the beginning of the Postglacial led very quickly to the domestication of plants and animals and the beginnings of agriculture. Those areas in western and eastern Asia, western Mexico, and South America became the centers of farming adaptations, the growing of crops and herding of animals, village occupation, and the use of pottery and ground stone tools. This story of the origins of agriculture is told in the next chapter.
p
DISCUSSION QUESTIONS 1. What is the evidence for a broad spectrum of food resources in the Postglacial? 2. Were plants important in past human diet?
For more review material and study questions, see the self-quizzes on your online learning center.
196
Page 196
Images and Ideas
w w w. m h h e . c o m / p r i c e i p 6 e
_ _
8:00 AM
3. Why was the dog the first domesticated animal? 4. Why are coasts and rivers important places of settlement for Postglacial hunter-gatherers? 5. Are there still groups of people living as hunter-gatherers?
Chapter 5 Postglacial Foragers
pri31057_ch05.pp164-197.qxd
z
10/1/09
8:00 AM
Page 197
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e.
Bailey, G., and P. Spikins, eds. 2008. Mesolithic Europe. Cambridge: Cambridge University Press. Binford, L. R. 2001. Constructing frames of reference: An analytical method for archaeological theory building using hunter-gatherer and environmental data sets. Berkeley and Los Angeles: University of California Press. The search for pattern in a massive study of ethnographically known hunter-gatherers. Burenhult, G. 1993. People of the Stone Age: Hunter-gatherers and early farmers. San Francisco: HarperCollins. A beautifully illustrated, popular book on the early Postglacial archaeology of foragers and farmers. Folkens, P. A., and T. D. White. 2000. Human osteology. Orlando, FL: Academic Press. A technically impressive introduction to the human skeleton. Habu, J. 2004. Ancient Jomon of Japan. Cambridge: Cambridge University Press. A survey of new evidence for the Jomon cultures of Japan. Jefferies, R. 2009. Holocene hunter-gatherers of the Lower Ohio River Valley. Tuscaloosa: University of Alabama Press.
Jefferies, R. W. 1987. The archaeology of Carrier Mills. Carbondale: Southern Illinois University Press. A popular account of the excavations of the Archaic site of Carrier Mills, Illinois, and of the general archaeology of that area. Kelly, R. L. 1995. The foraging spectrum: Diversity in huntergatherer lifeways. Washington, DC: Smithsonian Institution Press. An archaeologist’s look at hunter-gatherers. Larsen, C. S. 1999. Bioarchaeology: Interpreting behavior from the human skeleton. Cambridge: Cambridge University Press. The skeleton as a storehouse of information about the past. Lee, R. B., and R. Daly. 1999. Cambridge encyclopedia of hunters and gatherers. Cambridge: Cambridge University Press. Price, T. D., and J. A. Brown, eds. 1985. Prehistoric huntergatherers. Orlando, FL: Academic Press. A set of papers on early Holocene foragers in a variety of places around the world.
_ Suggested Readings
197
pri31057_ch06.pp198-265.qxd
10/1/09
C
8:02 AM
H
Page 198
A
P
6 T
E
R
S
I
X
Figure 6.1 A drawing from the Codex Florentino that shows an Aztec farmer planting maize seeds using a digging stick. The Spanish conquerors drew a number of these colorful codices to record views of life among the Aztecs.
_ _
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 199
The Origins of Agriculture Introduction The First Farmers The major transition in the course of human prehistory
C
orn is one of the most remarkable domesticated plants on earth, an important staple food for people and animals everywhere. The drawing in Figure 6.1 comes from the Codex Florentino, views of life among the Aztec made by the Spanish conquerors (more of the Codex can be seen in Figure 6.6). The great diversity of modern maize is an indication of its tremendous adaptability and genetic plasticity; this amazing plant grows under a variety of environmental and topographic conditions. Maize is one of the many species of plants and animals that humans have tamed over the past 10,000 years. As we have noted, our species spent the vast majority of history as huntergatherers. Our ancestry as food collectors, consuming the wild products of the earth, extends back more than 4 million years. Nevertheless, at the end of the Pleistocene, some human groups began to produce food rather than collect it, to domesticate and control wild plants and animals, achieving what is perhaps the most remarkable transition of our entire human past. Agriculture is a way of obtaining food that involves domesticated plants and animals. But the transition to farming is much more than simple herding and cultivation. It also entails major, long-term changes in the structure and organization of the societies that adopt this new way of life, as well as a totally new relationship with the environment. Whereas hunter-gatherers largely live off the land in an extensive fashion, generally exploiting a diversity of resources over a broad area, farmers intensively utilize a smaller portion of the landscape and create a milieu that suits their needs. With the transition to agriculture, humans began to truly master their environment. That a species is exploited intensively by humans does not automatically mean it will become domesticated. Although oak trees have supplied acorns for humans for thousands of years, they have not been domesticated. Domestication changes the physical characteristics of the plant or animal involved. The domestication process involves both the inherent characteristics of the plant or animal species (generational length, life cycle, plasticity) and the intensity and nature of the human manipulation. Agriculture requires several principal practices for long-term success: (1) propagation, the selection and sowing of seeds or breeding of animals; (2) husbandry, the tending of plants or animals during the growth period; (3) the harvesting of plants when ripe or the slaughter of animals at appropriate times; and (4) the storage of seeds and maintenance of animals through their nonproductive periods to ensure
199
< ( : < V : y < p : W
CHAPTER OUTLINE
Introduction The First Farmers 199
CONCEPT Explaining the Origins of Agriculture 205 SITE ’Ain Mallaha
209
CONCEPT Wheat, Barley, Pigs, Goats, and Sheep 211 SITE Abu Hureyra
214
CONCEPT Archaeobotany 218 SITE Jericho
221
CONCEPT Archaeozoology SITE Çatalhöyük
227
SITE Mehrgarh
232
CONCEPT Pottery SITE Ban-po-ts’un
234 236
CONCEPT Rice 239 SITE Khok Phanom Di
240
SITE Guilá Naquitz Cave CONCEPT Zea mays SITE Tehuacán
250
SITE Guitarrero Cave
224
246
243
254
CONCEPT Agriculture in Native North America 258 CONCEPT Breast-Feeding and Birth Spacing 260
Images and Ideas The Spread of Agriculture 262
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Figure 6.2 Harvesting and processing a cereal crop: scenes depicted on the walls of an Egyptian tomb.
w w w. m h h e . c o m / p r i c e i p 6 e
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
domestication The taming of wild plants and animals by humans. cultivation The human manipulation
_ _
or fostering of a plant species (often wild) to enhance or ensure production.
200
Page 200
annual reproduction. Plant propagation and husbandry involve cultivation— clearing fields, preparing the soil, weeding, protecting the plants from animals, and providing water. The evidence for early domesticated plants focuses on seed crops (Figure 6.2). The best-known early domesticates are the cereals—the grasses that produce large, hard-shelled seeds, nutritious kernels of carbohydrate that can be stored for long periods. The hard cereal grains, and occasionally the stems of these plants, were often burned during preparation or cooking in the past and thereby preserved to the present. Root crops are not well documented in the archaeological record because they lack hard parts that are more resistant to decay. Because they reproduce asexually from shoots or cuttings, it is difficult to distinguish domesticated varieties from their wild ancestors. Asexually reproducing plants may maintain exactly the same genetic structure through many generations, because a piece of the parent plant is used to start the daughter. Such plants may also exhibit great variation within a species, making domestication difficult to document. Root crops such as potatoes, yams, manioc, and taro may have been domesticated quite early. Archaeologists have started to identify them from prehistoric sites only recently. Animals were apparently domesticated initially for meat, with the exception of the dog. Dogs were tamed from wolves very early, perhaps 14,000 years ago in the Old World, and used for hunting and as pets, or even for food. Subsequently, however, several other animal species were domesticated
Chapter 6 The Origins of Agriculture
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 201
and herded for food and/or kept as beasts of burden. The animals domesticated earliest were pigs, goats, sheep, and cattle. The secondary products (such as milk, wool, horn, and leather) of these, and other, domesticated species also became important, as did their function as beasts of burden. The domestication of both plants and animals may be related to the storage of food. Such cereals as wheat, barley, corn, and rice have hard outer coverings that protect the nutritious kernel for some months, permitting the seed to survive until the growing season and offering very good possibilities for storage. Meat can be stored in the form of living, tame animals that are always available for slaughter. As such, storage provides a means to regulate the availability of food and to accumulate surplus. Questions concerning the origins of agriculture must focus on primary centers, where individual species of plants and/or animals were first domesticated (Figure 6.3). Secondary areas of agricultural development received plants and animals from elsewhere, although in many of those regions, some local plants and animals also were domesticated and used along with the introduced varieties. Primary centers for domestication were in Southwest Asia (Figure 6.4), East Asia, sub-Saharan Figure 6.3 The first appearance of domesticates by region and time.
Years B.C.
Southwest Asia
Africa
East Asia
Mesoamerica
0
South America
White potatoes
Lamb’s quarters
1000 Oil palm 2000 Sorghum, wheat 3000
North America
Camels
Maize, guinea pigs
Chickens (south-central Asia)
Marsh elder Goosefoot
Yams Gourds, potatoes, quinoa, manioc
Donkeys
Squash
4000 Maize, beans, chilis
Sheep, goats, cattle, barley, millet
5000
Cotton
Cattle, millet 6000
Llama, alpaca, peanuts Pigs, rice
7000
8000
9000
Gourds Goats, wheat, barley, cats, pigs, cattle, lentils Sheep Rye
Squash
Chilis, beans, squash
Figs Gourds
Introduction The First Farmers
_
201
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 202
Figure 6.4 Traditional techniques for harvesting and threshing grain in Southwest Asia have not changed for thousands of years.
llama A woolly South American ruminant camelid, used as a beast of burden. alpaca A domesticated South American herbivore with long, soft wool. camelid A ruminant mammal—such
_ _
as camel, llama, and extinct related forms—having long legs and two toes.
202
Africa, Mesoamerica, South America, and North America (Figure 6.5). The earliest known domesticates—wheat, barley, rye, peas, lentils, figs, pigs, goats, sheep, and cattle—appeared in the Old World—in Southwest Asia, between the eastern Mediterranean Sea and Afghanistan—at the end of the Pleistocene. Many other plants and animals—such as bread wheat, olives, grapes, and flax—were gradually added to this list. The origins of agriculture in Southwest Asia are discussed in detail in this chapter because the archaeology is well known and the process of domestication may have taken place there somewhat earlier than it did elsewhere. Agriculture also was invented in East Asia, perhaps in two or three different areas, sometime before 6000 B.C. Millet was first cultivated and pigs were domesticated in North China in villages dating to roughly 6000 B.C. Rice was initially cultivated in South China, possibly as early as the eighth or ninth millennium B.C., and somewhat later in Southeast Asia, around 4000 B.C. In all probability, root crops were under cultivation in that area, along with rice, sometime between 7000 and 3000 B.C. As a result of continuing research, we can expect the dates for all of East Asia to be pushed back somewhat earlier. Plants such as African rice, sorghum, and pearl millet were domesticated in sub-Saharan Africa after 2000 B.C. Cattle and goat herding was practiced in that area, where the new domesticates appeared. In the Americas, agriculture first developed in Mexico, in northwestern South America, and in eastern North America. In Mexico, gourds and squash were cultivated during the eighth millennium B.C.; avocados, chili peppers, beans, and possibly corn were cultivated later, around 5000 B.C. These crops provide all the essential nutrients for a healthy diet, and meat protein may not have been necessary. Domesticated animals never constituted an important part of the diet in this area, although turkeys, dogs, and the stingless honeybee were domesticated. Sites in the highlands of Peru contain evidence for the early domestication of gourds, tomatoes, beans, and potatoes by 3000 B.C. Some of these plants may have reached the mountains from an original habitat in the lowland jungles, but little is known about the prehistory of the Amazon Basin and other tropical areas of South America. Potatoes certainly were an indigenous crop; hundreds of varieties of wild and domesticated potatoes grow in this area today. In addition to plants, several animals were domesticated. The guinea pig was used for food, and the llama was probably domesticated and used to transport goods around the mountains of South America. The alpaca, a camelid like the llama, may have been domesticated for both meat and wool.
Chapter 6 The Origins of Agriculture
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 203
Abu Hureyra EASTERN WOODLANDS
Catalhöyük ’Ain Mallaha Jericho
Tehuacán
YELLOW RIVER
FERTILE CRESCENT
Ban-po-ts’un
Mehrgarh
MESOAMERICA
PACIFIC Khok Phanom Di O C E A N
SAHEL
Guilá Naquitz Cave
YANGTZE RIVER
Equator
NEOTROPICS Guitarrero Cave CENTRAL ANDES PACIFIC OCEAN
INDIAN OCEAN
ATLANTIC OCEAN
Sites discussed in the chapter
Abu Hureyra ’Ain Mallaha Jericho Çatalhöyük site at its largest extent
Mehrgarh Ban-po-ts’un Khok Phanom Di Guilá Naquitz Cave
episode of occupation
Tehuacán Guitarrero Cave
12,000
10,000
8000
6000
4000
2000
0
Years B.C.
Figure 6.5 Location of and timeline for primary centers of domestication.
_ Introduction The First Farmers
203
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Figure 6.6 Images of maize production from the Codex Florentino.
Page 204
In addition to those major centers of domestication, other areas witnessed the beginnings of horticulture or agriculture shortly after the end of the Pleistocene. Recent evidence suggests that agriculture was practiced very early in New Guinea. Radiocarbon dates from a digging stick found in what appear to be ancient agricultural fields indicate a date of around 7000 B.C. These fields were probably used for fruit and tuber crops such as yams. In eastern North America, several local plants such as marsh elder and goosefoot (Chenopodium) were domesticated by 1500 B.C., long before the introduction of corn from Mexico (see also “Agriculture in Native North America,” p. 258). Clearly, there was a trend toward domestication and agriculture on a global scale at the beginning of the Holocene. Following a discussion of various explanations for the origins of agriculture, this chapter traces the beginnings of farming in the different centers that have been identified. Because of the better quantity and higher quality of archaeological information from Southwest Asia, much of the discussion focuses on that area. Many geographic terms have been used to designate the area, including Near East, Middle East, and Southwest Asia. Near East refers to the Arabic countries of North Africa and southwestern Asia. The Middle East and Southwest Asia have similar boundaries, but the term Middle East reflects the view from Europe. Hence, Southwest Asia is the best way to describe the region. In Southwest Asia, we examine one site from before the transition to agriculture—’Ain Mallaha—and two early Neolithic communities—Abu Hureyra and Jericho—to see the changes that took place. Çatalhöyük, an enormous early Neolithic settlement, documents the consequences of the Neolithic revolution in terms of completely new ways of inhabiting the world. From Southwest Asia, the tour goes to South Asia and the site of Mehrgarh, an early Neolithic community in Pakistan. In East Asia, the sites of Ban-po-ts’un in northern China and Khok Phanom Di in coastal Thailand provide some sense of the Neolithic in that part of the world. There are at least three primary centers of domestication in the New World as well, in Mesoamerica, South America, and North America. The important early sites of Guilá Naquitz and in the Tehuacán Valley provide evidence of early plant domestication in Mexico. Excavations at Guitarrero Cave, high in the Andes, give us a glimpse of the process of domestication in South America. The summary section, “Images and Ideas,” considers the spread of agriculture from those primary centers, like the ripples spreading from a pebble thrown into a pond, to areas where domesticates were introduced from other places. Subsequent chapters explore the expansion of agriculture into Europe and parts of North America. We also discuss how this major change in human subsistence revolutionized economies, social organization, settlement, and ideology. Human society was never again the same after the beginnings of domestication. Even societies that continued to hunt and gather after the Neolithic were dramatically, and often drastically, affected by neighboring farmers.
_ _
204
Chapter 6 The Origins of Agriculture
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 205
Concept
<
Explaining the Origins of Agriculture The how and why of farming 400
350
mol mol –1
300
Atmospheric CO2,
It is remarkable that the process of domesticating plants and animals appears to have taken place separately and independently in a number of areas at about the same time. Read it again: The almost simultaneous appearance of domesticated plants and animals around the globe between roughly 10,000 and 5000 years ago is astounding. Given the long prehistory of our species, why should the transition to agriculture happen within such a brief period, a few thousand years in a span of over 6 million years of human existence? An important and dramatic shift in the trajectory of cultural evolution demands explanation. But such answers are hard to find. Views on and evidence for the origins of agriculture continue to be revised and updated. We can best understand ideas about the origins of agriculture from a historical perspective, considering the early theories first. Hypothetical explanations of why domestication occurred include the oasis hypothesis, the natural habitat hypothesis, the population pressure hypothesis, the edge hypothesis, and the social hypothesis. A consideration of these ideas also reveals much about the nature of archaeology and archaeologists. Theories about the origins of agriculture have often focused on the earliest evidence from Southwest Asia and, for that reason, may not be appropriate to all places where early domestication actually occurred. During the first half of the twentieth century, the best information on early farming villages came from riverine areas or oases with springs in North Africa and Southwest Asia—along the Nile River in Egypt or at Jericho in the Jordan Valley, for example. At that time, the end of the Pleistocene was thought to have been a period of increasing warmth and dryness in the earth’s climate. Researchers reasoned
250
33% rise
200
150 Intensive cultivation 100
50
0
35
30
25
20
15
10
5
0
Thousands of years before present
that because the ice ages were cold and wet, they should have ended with higher temperatures and less precipitation. Given that view of past climate, logic suggested that areas such as Southwest Asia, a dry region to begin with, would have witnessed a period of aridity at the end of the Pleistocene when vegetation grew only around limited water sources. The oasis hypothesis suggested a circumstance in which plants, animals, and humans would have clustered in confined areas near water. V. Gordon Childe, one proponent of this idea, argued that the only solution to the competition for food in these situations would have been for humans to domesticate and control the animals and the plants. In this sense, domestication emerged as a symbiotic relationship for the purpose of human survival.
Figure 6.7 A graph of carbon dioxide levels in the atmosphere for the past 35,000 years. Note particularly the big increase between 15,000 and 10,000 years ago and then again in recent decades.
oasis hypothesis The theory that domestication began as a symbiotic relationship between humans, plants, and animals at oases during the desiccation of Southwest Asia at the end of the Pleistocene.
Concept Explaining the Origins of Agriculture
205
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
The conditions of incipient desiccation . . . would provide the stimulus towards the adoption of a food-producing economy. Enforced concentration by the banks of streams and shrinking springs would entail an intensive search for means of nourishment. Animals and men would be herded together in oases that were becoming increasingly isolated by desert tracts. Such enforced juxtaposition might promote that sort of symbiosis between man and beast implied by the word domestication. —V. Gordon Childe (1951)
The food-producing revolution seems to have occurred as the culmination of the ever increasing cultural differentiation and specialization of human communities. Around 8000 B.C., the inhabitants of the Fertile Crescent had come to know their habitat so well that they were beginning to domesticate the plants and animals they had been hunting and gathering. —Robert Braidwood (1960)
natural habitat hypothesis The
theory that the earliest domesticates appeared in the area that their wild ancestors inhabited. population pressure hypothesis
The theory that population increase in Southwest Asia upset the balance between people and food, forcing people to turn to agriculture as a way to produce more food. edge hypothesis The theory that the
_ _
need for more food was initially felt at the margins of the natural habitat of the ancestors of domesticated plants and animals; a revised version of the population pressure hypothesis.
206
Page 206
More recently, detailed information on climate change has come from a most unlikely place—the glaciers of Greenland. Deep corings of the ice sheets there have provided a layered record of changes in temperature and other aspects of climate for the past 100,000 years and more. One of the very interesting results of this research was the documentation of a 33% increase in atmospheric carbon dioxide at the end of the Pleistocene (Sage, 1995) (Figure 6.7). Higher levels of CO2 would foster the expansion of temperate species such as grasses, which include many of the ancestors of the major domesticated species. The full implications of such changes in the atmosphere are not yet clear, but the changes may have played a role in the transition from hunting to farming. During the 1940s and 1950s, however, new evidence indicated that there had been no major climate changes in Southwest Asia at the close of the Pleistocene—no crisis during which life would have concentrated at oases. The new information forced a reconsideration of the origins of agriculture. The late Robert Braidwood pointed out—in his natural habitat hypothesis—that the earliest domesticates therefore should appear where their wild ancestors lived. That area, the “hilly flanks” of the Fertile Crescent in Southwest Asia, should be the focus of investigations. Braidwood and a large team of researchers excavated at the site of Jarmo in northern Iraq. The evidence from this early farming village supported his hypothesis that domestication did indeed begin in the natural habitat. Braidwood did not offer a specific reason as to why domestication occurred, other than to point out that technology and culture were ready by the end of the Pleistocene, that humans were familiar with the species that were to be domesticated. At that time, archaeologists and others considered farming to be a highly desirable and welcome invention, providing security and leisure time for prehistoric peoples. Once human societies had recognized the possibilities of domestication, they would have immediately started farming.
Chapter 6 The Origins of Agriculture
Lewis Binford, of Southern Methodist University, challenged those ideas in the 1960s and proposed the population pressure hypothesis. Binford argued that farming was backbreaking, time-consuming, and labor-intensive. Citing studies of living hunter-gatherers, he pointed out that they spent only a few hours a day obtaining food; the rest of their time was for visiting, talking, gambling, and otherwise enjoying life. Even in very marginal areas, such as the Kalahari Desert of South Africa, food collecting is a successful adaptation, and people rarely starve. Binford argued, therefore, that human groups would not have become farmers unless they had no other choice, that the origin of agriculture was not a fortuitous discovery but a last resort. Binford made his point by positing an equilibrium between people and food, a balance that could be upset by either a decline in available food or an increase in the number of people. Since climatic and environmental changes appeared to be minimal in Southwest Asia, Binford thought it must have been increased population size that upset the balance. Population pressure was thus introduced as a causal agent for the origins of agriculture: More people required more food. The best solution to the problem was domestication, which provided a higher yield of food per acre of land. At the same time, however, agricultural intensification required more labor to extract the food. Binford further suggested that the effects of population pressure would have been felt most strongly not in the core of the natural habitat zone, where dense stands of wild wheat and large herds of wild sheep and goats were available, but at the margins, where wild foods were less abundant. This theory, incorporating ideas about population pressure and the margins of the Fertile Crescent, has become known as the edge hypothesis. Binford’s concern with population was elaborated by Mark Cohen, of the State University of New York– Plattsburgh. Cohen argued for an inherent tendency for growth in human population, a pattern responsible for
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
the initial spread of the human species out of Africa, the colonization of Asia and Europe, and eventually colonization of the Americas as well. After about 10,000 B.C., according to Cohen, all the habitable areas of the planet were occupied, and population continued to grow. At that time, there was an increase in the use of less desirable resources in many areas. Land snails, shellfish, birds, and many new plant species were added to the human diet around the end of the Pleistocene. Cohen argued that the only way for a very successful, but rapidly increasing, species to cope with declining resources was for them to begin to cultivate the land and domesticate its inhabitants, rather than simply to collect the wild produce. Domestication for Cohen was a solution to problems of overpopulation on a global scale. Others, arguing that the transition to farming and food storage and surplus cannot be understood simply in terms of environment and population, have developed social hypotheses to explain the origins of agriculture. Barbara Bender, of the University of London, and Brian Hayden, of Simon Fraser University, for example, have suggested that the success of food production may lie more in the ability of certain individuals to accumulate a surplus of food and to transform that surplus into more valued items, such as rare stones and metals. From this perspective, agriculture was the means by which social inequality emerged and egalitarian societies became hierarchical. There are several other useful theories about why human societies adopted agriculture at the end of the Pleistocene. Geographer Carl Sauer suggested that agriculture began in the hilly tropics of Southeast Asia, where sedentary groups with knowledge of the rich plant life of the forest might have domesticated plants for poisons and fibers. Botanist David Rindos has argued that domestication was a process of interaction between humans and plants, evolving together into a more beneficial symbiotic relationship. In a fascinating book titled Birth of the Gods and the Origins of Agriculture,
Page 207
Change in the demographic structure of a region which brings about the impingement of one group on the territory of another would also upset an established equilibrium system, and might serve to increase the population density of a region beyond the carrying capacity of the natural environment. Under these conditions, manipulation of the natural environment in order to increase its productivity would be highly advantageous. —Lewis Binford (1968)
Figure 6.8 Two statues of plaster found with several others at the site of ’Ain Ghazal in Jordan. The eyes are cowrie shells set in bitumen. The taller statue is 90 cm (about 3 ft) high. Such figures likely reflect changing religious beliefs in the Neolithic.
French archaeologist Jacques Cauvin argues that the important changes associated with the “Neolithic revolution” were more cultural than economic. That is, the transition to farming involved concepts and ideas as much as or more than cultivating and herding. Specifically, he suggested that domestication was preceded by the emergence of new religious practices and symbolic behavior (Figure 6.8). The transformation of hunter-gatherers that allowed them to view their habitat in a different way also promoted the more active exploitation of that environment. Some problems with all these theories can be seen in a brief consideration of the evidence from Southwest Asia. The earliest agricultural villages, places such as Abu Hureyra and Jericho, were indeed located at the margins of the natural habitat. Attempts to artificially reproduce stands of wild wheat there may have resulted in
Technology and demography have been given too much importance in the explanation of agricultural origins; social structure too little. . . . Food production is a question of techniques; agriculture is a question of commitment. . . . Commitment is not primarily a question of technology but of changing social relations. This account has chosen to emphasize the social properties of gatherer-hunter systems; to show how alliance structures, and the individuals operating within these structures, make demands on the economic productivity of the system; how demography and technology are products of social structure rather than independent variables. —Barbara Bender (1978)
social hypothesis The theory that
domestication allowed certain individuals to accumulate food surplus and to transform those foods into more valued items, such as rare stones or metals, and even social alliances.
Concept Explaining the Origins of Agriculture
207
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 208
Non Sequitur by Wiley
NON SEQUITUR © 1995 Wiley Miller. Dist. By Universal Press Syndicate. Reprinted with permission. All rights reserved.
By approximately 11,000 or 12,000 years ago, hunters and gatherers, living on a limited range of preferred foods, had by natural population increase and concomitant territorial expansion fully occupied those portions of the globe which could support their lifestyle with reasonable ease. By that time, in fact, they had already found it necessary in many areas to broaden the range of wild resources used for food in order to feed growing populations. I suggest that after that time, with territorial expansion becoming increasingly difficult and unattractive as a means of adjusting to growing population, they were forced to eat more and more unpalatable foods, and in particular to concentrate on foods of low trophic level and high density. —Mark Cohen (1977)
y _ _
Were human groups forced to become farmers, or was it a decision they made?
sedentism Living in permanent, yearround contexts, such as villages.
208
domestication. However, human populations were not particularly large just before agriculture. Several sites show signs of abandonment in the levels beneath those layers that contain the first domesticated plants. The most recent climatic evidence indicates that there was, in fact, a period of slightly cooler and moister temperatures in Southwest Asia at the end of the Pleistocene, which may have greatly expanded the geographic range of wild wheats and barley, making them available to more human groups and fostering the process of domestication. In combination with the evidence for changes in CO2 at that time, the possibility becomes intriguing. At the same time, however, these species were present in North Africa and parts of Southwest Asia during the Pleistocene and were not domesticated. Other factors were at work. Some theories may seem reasonable in one of the primary centers of domestication but not in another. The sequence of events in two areas is of particular interest here. In Southwest Asia, permanent settlements are known from 11,000 B.C., before the presence of direct evidence for domesticated plants or animals. Cultivated plants appeared about 9000 B.C.; animals were probably not herded until perhaps 8500 B.C. Pottery did not come into general use until around 7500 B.C. In Mesoamerica, however, the archaeological sequence reveals that domesticated plants first appeared around 5000 B . C ., followed by pottery and
Chapter 6 The Origins of Agriculture
then permanent villages several thousand years later. Domesticated animals were never important in this area. The differences in these two areas indicate that sedentism and cultivation are not totally dependent on each other. It is also clear that domesticated animals are not a part of the equation in all areas. Because of the difficulties in trying to excavate phenomena such as social relations and population pressure, many of the current theories are hard to evaluate. Any adequate explanation of the agricultural transformation should deal not only with how it all began but also with why it happened rather suddenly. Population and climatic change certainly play a role in cultural evolution, but we cannot yet say precisely why plants and animals began to be domesticated shortly after the end of the Pleistocene. The how and the why of the Neolithic transition remain among the more intriguing questions in human prehistory. Simply put, there is, as yet, no single accepted general theory for the origins of agriculture. No common pattern of development is apparent in the various areas where domestication first took place. At the same time, of course, the evidence we have is still scanty and limited. This chapter examines the origins of agriculture in more detail in the several primary centers where it first appeared: Southwest Asia, East Asia, Mexico, and South America, as well as Africa and eastern North America.
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 209
Site
:
’Ain Mallaha Pre-Neolithic developments in Southwest Asia
Discussions about the origins of agriculture often focus on Southwest Asia, for several reasons: (1) The earliest evidence for plant domestication from anywhere in the world is found here, (2) there is a reasonable amount of information available from excavations and other studies, and (3) Southwest Asia is often considered the “cradle of Western civilization.” To better understand the origins of agriculture in this region, it is useful to look at human settlements that preceded domesticated plants and animals. The period just before agriculture, roughly 11,000–9000 B.C., is referred to as the Natufian. Most of the evidence for this period comes from the Levant, a mountainous region in the eastern Mediterranean. The period was characterized by an increase in the number of sites, and therefore people, coinciding with a period of more rainfall and abundant vegetation. The natural habitat was rich in wild plants and animals, resources that supported permanently settled communities before any evidence of domestication. The Natufian site of ’Ain Mallaha (ein ma-LA-ha) lies beside a natural spring on a hillside overlooking the swamps of Lake Huleh in the upper Jordan Valley of Israel (see Figure 6.12). ’Ain Mallaha was one of the earliest villages anywhere in the world, dating to 11,000–9000 B.C. The entire settlement covered an open area of about 2000 m2 (1⁄2 acre, the size of a large hockey rink), with a population estimated at 200– 300 people. Excavations between 1955 and 1973 by Jean Perrot, of the French Archaeological Mission in Jerusalem, uncovered three successive layers with the remains of permanent villages. Each layer contained a number of round houses, ranging from 3 to 8 m (10 to 25 ft) in diameter (Figure 6.9). House entrances faced downhill toward the water.
The remarkable architecture consists of large substantial houses with stone foundations standing to a height of almost 1 m (3 ft). Wooden center posts may have supported conical roofs. Stone-lined square or oval hearths and bins were found in the center of the rooms or against the walls. Mortars and querns—grinding tools and surfaces for preparing grain—were occasionally set into the floor (Figure 6.10). Although the structures were built close together, the community had a centrally located open area with round storage pits. The ground stone artifacts include plates, bowls, mortars, and pestles, indicating a need for containers at this time (Figure 6.11). Several objects are decorated with elaborate geometric designs. Carved limestone figurines of a human body, a human face, and a tortoise also were found. The flaked stone industry is rich, with more than 50,000 pieces. The bone tools include awls, skewers, needles, and fishhooks. The animal bones found at the site come from wild pig, three kinds of deer, wild goat, wild cattle, wild horse, and
Figure 6.9 Excavations at ’Ain Mallaha exposing burials under the house floors.
Levant A mountainous region paral-
leling the eastern shore of the Mediterranean, including parts of the countries of Turkey, Syria, Lebanon, and Israel. mortar A bowl-shaped grinding tool, used with a wood or stone pestle for grinding various materials. quern A stone grinding surface for
preparing grains and other plant foods and for grinding other materials. gazelle One of several species of small to medium swift and graceful antelopes native to Asia and Africa. net-sinker A small weight attached to fishing nets.
Site ’Ain Mallaha
209
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 210
Multiple burial in pit
Figure 6.10 Two circular houses at ’Ain Mallaha. These structures have rock wall foundations and often contain grinding equipment, storage bins, and pits. Burials were often placed in abandoned storage pits.
N
Entrance?
Bin
Pestle Mortar
Hearth
Pestles
0
Figure 6.11 Various artifacts from ’Ain Mallaha. a–j: chipped stone tools; k–m, o: ground stone tools; n: a mortar; p–r: ground stone containers. (c)
(b)
(d)
(e)
(f)
(h)
(g)
(a)
(j)
(k)
(m) (l)
(i)
(p)
(o)
(n)
(r) (q)
sickle A tool for cutting the stalks of
_ _
cereals, especially wheat. Prehistoric sickles were usually stone blades set in a wood or antler handle.
210
Chapter 6 The Origins of Agriculture
1m
gazelle. Gazelle is the most common game animal at sites of this period. Bird, fish, tortoise, and shellfish remains also were found. The lake was clearly an important resource for these people, as indicated by the fish and shellfish remains, along with net-sinkers. The high incidence of decay in the teeth of individuals buried at ’Ain Mallaha suggests that carbohydrates from cereals or other plants were consumed in quantity. Wild barley and almonds were found charred in excavations, and it is clear from the abundance of sickle blades and other plant-processing equipment that wild cereals played an important role in the diet. Two kinds of burials were found at ’Ain Mallaha: (1) individual interments, including child and infant burials beneath stone slabs under the house floors; and (2) collective burials in pits, either intact or as secondary reburials after soft tissue had disappeared. Most of the 89 graves were found outside the houses. Abandoned storage pits were often reused for burial purposes. Many graves contained red ochre, and limestone slabs covered several of the simple graves. Four horns from gazelle were found in one grave, and in another, an old woman was buried with a puppy. Shells from the Mediterranean and rare greenstone beads or pendants from Syria or Jordan were occasionally placed with the burials, but grave goods were generally rare.
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 211
Concept
<
Wheat, Barley, Pigs, Goats, and Sheep The appearance of the first farmers in Southwest Asia at the end of the Pleistocene
Southwest Asia is a fascinating region. Perhaps too well known today for the political problems that beset it, the area also was the home of the earliest domesticated plants and animals, as well as some of the world’s first civilizations. Southwest Asia is an enormous triangle of land, approximately the size of the contiguous United States. The area is bounded on the west by Turkey and the Mediterranean, on the south by Saudi Arabia and the Indian Ocean, on the north by the Black and Caspian seas, and on the east by Afghanistan, at the edge of South Asia. Southwest Asia is a series of contrasts. Some of the highest and lowest places in the world are found there, along with both rain forest and arid desert. Snow-capped mountains are visible from scorching-hot wastelands. Water is an important resource; arable land with fertile soil is scarce. The environment of this area can be visualized as a series of bands, driest in the south and moistest in the north. Arabia is largely sand and desert; Mesopotamia, the classic region between the Tigris and Euphrates rivers, is too dry for farming unless some form of irrigation is used. Mesopotamia has nothing to do with the origins of agriculture. Most plants cannot survive in areas with less than 300 mm (1 ft) of rain each year. The line showing this 300-mm rainfall isobar stretches along an arc of mountains. The Zagros Mountains of western Iran, the Taurus Mountains of southern Turkey, and the highlands of the Levant along the eastern Mediterranean shore form a region where more rain falls and a variety of plants grow in abundance. The area is known as the Fertile Crescent, a name that reflects the variety of plants and animals that became the basic staples of many agricultural societies (Figure
6.12). This region is the natural habitat of many of the wild ancestors of the first species of plants and animals to be domesticated at the end of the Pleistocene—the wild wheats and barleys; the wild legumes; and the wild sheep, goats, pigs, and cattle that began to be exploited in large numbers at the time of the first agriculture. Some 20,000 years ago, a series of developments began in Southwest Asia that set the stage for village farming. Climatic conditions during this period are not completely understood, but some general patterns are known. Around 18,000 B.C., global temperatures were about 6°C (10°F) cooler than they are today. A warming trend began about 14,000 B.C. and increased to a maximum temperature around 4000 B.C. Climate at the very beginning of the Neolithic, 11,000 years ago, was somewhat variable. Precipitation changes were not dramatic, but in an arid area, minor changes in rainfall can have a significant impact on vegetation. Rainfall was lowest during periods of maximum cooling around 18,000 B.C. As temperature and precipitation increased, the forest zone expanded in Southwest Asia, and the number of species was greater than it is today. After about 8000 B.C., however, continuing increases in temperature likely resulted in more evaporation, so that effective precipitation began to decline and the forest cover shrank. Within this climatic and environmental context, a gradual change from a broad-spectrum diet, focusing on the many wild species of the region, to a diet that concentrated on a few domesticated plants and animals can be seen. In the late Paleolithic, after 20,000 years ago, groups of hunter-gatherers lived in small, seasonal camps throughout the area. Although they exploited a range of resources, they focused on
Fertile Crescent An upland zone in Southwest Asia that runs from the Levant to the Zagros Mountains.
Concept Wheat, Barley, Pigs, Goats, and Sheep
211
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 212
Figure 6.12 Locations of Southwest Asian sites mentioned in the chapter. The shaded area marks the Fertile Crescent.
ARMENIA
TURKEY
AZERBAIJAN
KAZAKHSTAN
Hallam Cemi Caspian Sea
Catalhöyük Abu Hureyra
CYPRUS Mediterranean Sea
LEBANON
ISRAEL
Shanidar
SYRIA
’Ain Mallaha Jericho
IRAN
IRAQ
JORDAN EGYPT
KUWAIT SAUDI ARABIA 0
animals such as the gazelle. Plant foods are not common in the sites from this period. In the period just preceding the Neolithic, there was more intense utilization of plant foods. Particularly noticeable is the range of equipment for processing plants: sickle blades and grinding stones, along with storage pits and roasting areas for preparing wild wheat. Sites were often located in areas of cultivable land, but such settlements depended on wild cereals, as evidenced in the remains of wild wheat and barley. These same locations were occupied during the Neolithic, too, probably because of the quantity or quality of arable land. Hunting continued, and more immature animals were killed, including gazelles and wild goats. Between 9000 and 8000 B.C., changes in the size, shape, and structure of several cereals indicate that they had been domesticated. The archaeological data from Jericho and Abu Hureyra (A-boo hoo-RAY-rah), for example, mark this transition. The Neolithic, defined by the appearance of domesticated plants, began at that time. The earliest known domesticated cereal, rye, has been dated to 10,000 B.C. at the site of Abu Hureyra in Syria. In fact, eight or nine
_ _
212
Chapter 6 The Origins of Agriculture
300 km
“founder” plants were domesticated during the period 9000–7000 B.C., including three cereals—emmer wheat, einkorn wheat, and barley—and four or five pulses—lentils, peas, bitter vetch, chickpeas, and maybe fava beans. (Pulses are the edible seeds of leguminous plants, such as peas and beans.) Flax also was domesticated during this period and probably was used for oil and fiber; linen is made from the fibers of the flax plant. The first evidence for domestication of these founder plants comes from the same areas in which their wild ancestral stock is common. For example, genetic analysis has identified the original homeland of einkorn’s (a primitive wheat) wild ancestor in southeastern Turkey. The archaeological evidence can tell us when and where, but not why and how. The transition to the Neolithic was marked not by abrupt changes but by increasing emphasis on patterns that appeared during the Natufian. The number and the size of prehistoric communities expanded greatly during the early Neolithic, as populations apparently concentrated in settlements. The first towns appeared. Major changes in human diet, and probably in the organization of society
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
as well, began to take place. Some of the first domesticated animals are from Hallam Çemi, a very early Neolithic site in eastern Turkey, dating to around 9000 B.C. Excavated by Michael Rosenberg, of the University of Delaware, the site was a village of small, round houses and one larger, nondomestic building with a centrally located feasting area. The food remains at the site include wild sheep and goats, along with various nuts and wild legumes. Wild cereals were not an important part of the diet. About 10% of the animal bones came from pigs. Evidence for the domestication of these pigs is seen in the sex and age of death of the animals. Most of the bones were from young female animals. It appears that the inhabitants were selecting suckling pigs to eat, supporting the argument that these animals were controlled, or herded. In addition, the teeth of the pigs at Hallam Çemi are smaller than those of their wild relatives. Some of the more interesting recent discoveries come from the island of Cyprus in the eastern Mediterranean. Cyprus lies more than 60 km (40 mi) from the coast of the Near East. Nevertheless, early Neolithic farmers had sailed here from the mainland by 8000 B.C. Even more remarkably, they brought a number of domestic and
Page 213
wild animal species with them to populate the island. Red deer were not present on the island of Cyprus prior to this time, strongly suggesting that the waterborne farmers brought this animal with them. The farmers also introduced cattle and sheep and goats along with a number of crop plants originally domesticated in the Near East. Finally, they appear to have brought their pets. The earliest known domestic cat skeleton has been found buried next to a human grave from a settlement dating to 7300 B.C. Cats may have been domesticated to control the mice that were attracted to the stored grain of the farming villages. By 7500 B.C., domesticated sheep and goats had made their first appearance in the Levant, and a number of changes in architecture had occurred. Pottery was invented in Southwest Asia around 7500 B.C. to serve as easily produced, waterproof containers. These dishes were probably used for holding liquids, for cooking a gruel made from wheat and barley (bread was a somewhat later invention), and for storing materials. The complete Neolithic package of domesticates, village architecture, and pottery was thus in place shortly before 7000 B.C., as the Neolithic revolution began to spread to Europe and Africa.
Was the adoption of agriculture almost instantaneous in human prehistory?
y
_ Concept Wheat, Barley, Pigs, Goats, and Sheep
213
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 214
Site
:
Abu Hureyra Hunter-gatherers and early farmers in northern Syria
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on Abu Hureyra, see the Internet exercises on your online learning center.
_ _
tell A mound composed of mud bricks and refuse, accumulated as a result of human activity.
214
In 1974, the site of Abu Hureyra in northern Syria was submerged beneath the waters behind a new dam on the Euphrates River. Fortunately, in 1972 and 1973, before the water level in the reservoir rose and flooded the area, rescue excavations uncovered parts of this site, one of the largest early Postglacial communities in Southwest Asia. Excavations were conducted by A. M. T. Moore, of the University of Rochester, and his colleagues. The tell—an accumulated mound of occupation debris—covered 11.5 ha (about 30 acres), with deposits from the Natufian and the early Neolithic up to 8 m (25 ft) high in some places. One million cubic meters (1.3 million cubic yards) of earth were removed during the excavations. The primary component of the tell was the decayed mud walls of the generations of houses that were built there, along with the artifacts and food remains left behind by the inhabitants. The layers indicated an uninterrupted occupation of the mound from approximately 10,500 B.C. to 6000 B.C., through the Natufian and the Neolithic periods in Southwest Asia. Abu Hureyra thus contains one of the best available records of the changes that took place as farming and herding first began (Figure 6.13). The mound lies at the edge of the Euphrates River, with the river floodplain on one side and dry, level steppe on the other. The area today receives approximately 200 mm (8 in) of rainfall per year; cultivation is difficult without irrigation. During the Natufian occupation of the site, however, the climate was warmer and wetter. An open forest of oak and pistachio trees grew on the steppes nearby, with dense stands of wild grasses among the trees. These grasses, probably no more than 1–2 km (about 1 mi) distant, included wild wheats, rye, and various pulses (lentils and legumes).
Chapter 6 The Origins of Agriculture
The Mesolithic settlement was located on the northern side of the tell, adjacent to the Euphrates River. The settlement may originally have been placed here along the migration route of the gazelle herds. These animals were killed in great numbers during the spring migration. The settlement consisted of small, circular pit dwellings dug into the original ground surface. These structures had a framework of wooden posts supporting the wall and roof. Almost 1 m (3 ft) of debris accumulated during this first phase of occupation, between approximately 10,500 and 9000 B.C. The population of the site is estimated to have been between 200 and 300 at that time. Clearly, the bulk of their food came from the wild plants, some of which were staples. The plant remains at the site indicate a year-round occupation in both the Mesolithic and the Neolithic periods. The excavators used sophisticated techniques at Abu Hureyra to recover more than 500 liters (140 gallons) of plant remains from the site. From the Natufian levels, there was evidence for wild lentils, hackberry fruit, caper berries, and nuts from the turpentine tree, related to pistachios. Most intriguing, however, were the remains of wild wheat, barley, and rye. Around 10,000 B.C., the climate became cooler and drier and the nearby stands of wild cereals and other plants retreated more than 100 km (62.5 mi) to the higher elevations of the Fertile Crescent. Fruits and seeds of droughtsensitive plants from an oak–pistachio open woodland disappeared at Abu Hureyra. Then wild lentils and other legumes declined. Local vegetation around the site appears to have changed from moist, woodland steppe to dry, treeless steppe. Wild wheats continued to be consumed at the site even though their habitat in the area had
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Abu Hureyra
Ghab
T Ri v igris
Mediterranean Sea
Zeribar
er
Mirabad
E
Huleh 0
300 km
te s hra up iv
R
Forest and fairly dense woodland (including montane forest, aumediterranean sclerophyllus woodland, and xeric deciduous oak-Rosaceae woodland). Oak-terebinth-Rosaceae woodland (a moasic of woodland and open areas dominated by annual grasses).
er
(a)
Terebinth-almond woodland steppe, involving a thin scatter of trees in what were otherwise grass-dominated steppe formations. Abu Hureyra
s Ri Tigri
Mediterranean Sea
Zeribar
er
v
Mirabad
E te s hra up iv
R
er
Areas (within the previous two zones) supporting extensive stands of wild wheats and ryes. Steppe, dominated by wormwoods, perennial chenopods, and perennial tussock-grasses. Mosaic of areas dominated by trees of montane forest, aumediterranean woodland, xeric deciduous woodland, and woodland steppe, most of them probably growing as relatively thin scatters.
(b)
Ghab
Abu Hureyra
Mediterranean Sea
sR Tigri
Zeribar er
iv
Mirabad
E
Huleh
te s hra up
iv
R
er
(c)
Abu Hureyra Zeribar er
iv
Mediterranean Sea
sR Tigri
The partial die-back zone, characterized by isolated pockets of trees with wild cereals and legumes (micro-refugia) that will have survived in moist hollows and at breaks in north-facing slopes, surrounded by areas littered with dead trees. The different densities of dots reflect the lower density of these scattered pockets toward the outer fringes of this zone. The zone of total arboreal dieback, characterized by dead trees, without any of the isolated pockets of living trees of the previous zone, barring terebinths and caper bushes growing in some wadi bottoms.
Mirabad
E te s hra up
iv
R
been eliminated. The excavators believe that the Natufian inhabitants practiced plant husbandry of wild cereals before changes in the glume and rachis brought about by domestication were evident. Significantly, the earliest known domesticated plant, rye, appeared at that time. Grinding stones and milling equipment also point to the importance of cereals in the diet during that period. Experiments by Gordon Hillman, of the University of London, were designed to estimate the amount of time needed for wild cereals to change to the domesticated variety through the process of cultivating and harvesting the wild seeds and replanting them. This study indicated that the domestication of the plants could have taken place within a period of less than 300 years, perhaps no more than 25 years. Shortly after the initial domestication of rye and the probable cultivation of wild wheats, lentils and legumes reappeared in the deposits and increased. By 8500 B.C., the range of domesticated plants included rye, lentils and large-seeded legumes, and domesticated wheats. Clearly, plant domestication began in the Natufian period at Abu Hureyra, perhaps in response to the disappearance of the wild stands of these important foods. Two tons of animal bone, antler, and shell also were removed during the excavations. Shells from river mussels, fish bones, and bone fishhooks indicate that the inhabitants obtained food from the Euphrates River, as well as from the surrounding hills. Gazelle bones dominate the lower layers at Abu Hureyra and constitute 80% of all animal bones from the Mesolithic and early Neolithic periods. By the beginning of the Neolithic, however, sheep and goats had been domesticated and were being herded. After 7500 B.C., the number of gazelle bones dropped sharply, and sheep and goats became much more important in the diet (Figure 6.14). During the subsequent phases of the Neolithic at Abu Hureyra, domesticated cattle and pigs were added to the larder. Abu Hureyra grew quickly to become the largest community of its day,
Page 215
er
(d)
Figure 6.13 The landscape of Southwest Asia at the end of the Pleistocene. These four maps show a sequence of changing vegetation for four periods of the late Pleistocene and early Holocene. Note particularly the changes at Abu Hureyra, where the vegetation goes from (a) grass steppe, to (b) lightly forested with wild wheats and ryes during an episode of cooler and wetter conditions around 12,000 years ago, and then reverts to (c, d) dry grass steppe. These changes may have important implications for the domestication of plants. The four named locations (Huleh, Ghab, Zeribar, and Mirabad) are sites where environmental data for this reconstruction were obtained.
_ Site Abu Hureyra
215
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 216
Abu Hureyra 1 Plants
Plants
Wild cereals
Parkwoodland
Steppe and woodlandsteppe
Abu Hureyra 2A Wheats and ryes
Wild lentils
Barley
Domestic rye (Abu Hureyra 1B and 1C)
Legumes
Terebinth nutlets
Wild plants—Euphrates Valley
Small-seeded grasses
Wild plants—steppe Animals Sheep and goat herding
Small-seeded legumes Feather grass seed
Gazelle hunting
Perennial chenopod seed
O N D J F M A M J J A S Season 1 Season 2 Season 3 Season 4
Roots and tubers of dryland plants Key Plants
Knotgrass seeds Valley bottom
Tillage, sowing and tending of crops Club-rush seeds
Harvesting and processing of crop Availability of wild plants
Wild millet seed
Estimated period of plant food use in storage
Club-rush tubers and bulrush rhizomes Animals
Gazelle
Potential extended storage period Animals Herding activity Interception of animal migrations Hunting of nonmigratory wild animals
Onager
Estimated period of stored meat use Wild sheep Hare A M J J A S O N D J F M Season 1 Season 2 Season 3
Figure 6.14 The seasonal availability and use of plants and animals at Abu Hureyra during the pre-Neolithic and the Neolithic. During Abu Hureyra 1, ca. 12,000 years ago, there were no domesticated plants or animals, and a wide variety of species contributed to the diet during all seasons of the year. In the Early Neolithic (Abu Hureyra 2A), ca. 10,500 years ago, domesticated plants and herded sheep and goats played an important role in subsistence. Wild plants remained important, and the seasonal hunting of wild gazelle continued to supplement the food provided by herds of domestic sheep and goats. In later periods, the wild species declined in importance.
with 2000–3000 inhabitants in an area of about 11.5 ha (30 acres). Houses were rectangular, with mud-brick walls that were plastered and whitewashed (Figure 6.15). Plaster also was used to make heavy rectangular containers. Clay was used for beads and figurines, but pottery was not present in this
_ _
216
Chapter 6 The Origins of Agriculture
level of the site. The importance of this community is documented by the quantity and variety of exotic materials that arrived there through trade and exchange: cowrie shells from the Mediterranean or Red Sea, turquoise from the Sinai Peninsula, and obsidian, malachite, agate, jadeite, and serpen-
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 217
TABLE 6.1 Changes in Environment, Economy, and Settlement Types at Abu Hureyra Years B.C.
Period
Environment
6000 6300 7500
Neolithic
Decline in gazelle 8500 9000
Intermediate
9500
10,000
10,500
Cooler and drier; retreat of forest; dry, open steppe
Natufian
Younger Dryas Open, rolling steppe and grassland with nearby park woodland of oak and pistachio, wild cereals and legumes
tine from the mountains of Turkey. These rare stones were made into large, thin “butterfly” beads, often found in burials. Table 6.1 summarizes the various changes at Abu Hureyra from 10,500 B.C. to 6000 B.C. By 6000 B.C., Abu Hureyra had been abandoned. A similar pattern is seen at other Neolithic sites in the Levant at that time. It seems likely that increasingly arid conditions reduced agricultural productivity and made herding a more viable enterprise. It may be at that time that nomadic herding became the dominant mode of life, much like that of the pastoralists who still roam parts of Southwest Asia with their herds of sheep and goats. The evidence from Abu Hureyra indicates that cultivation began in ancient Southwest Asia in a small, sedentary village of hunter-gatherers around 10,000 years ago during a period of environmental change. The disappearance of the habitat for wild species coincided with the early domestication of rye and eventual cultivation of wheat, lentils, and legumes. But the transition from dependence on wild, gathered foods and hunted animals to domesticated varieties took 2500 years. Not all families were initially involved in
Economy
Settlement
Mixed farming cereals, legumes, sheep, goats, cattle, pigs Cereal and legume cultivation, sheep and goat husbandry Cereal and legume cultivation, plant gathering, gazelle hunting, domesticated sheep and goats Cereal and legume cultivation, plant gathering, gazelle hunting Wild einkorn wheat out of habitat. Domestication of rye, plant gathering, gazelle hunting
7 ha clustered mud-brick houses > 16 ha clustered mudbrick houses 8 ha clustered mud-brick houses
Gathering wild plants, hunting gazelle
Pit dwellings
farming, and the number increased over time. The first sheep and goat husbandry appeared around 8500 B.C., followed by that of cattle and pigs. The general sequence involves settlement in villages, followed by plant cultivation and subsequent animal herding. This pattern of a gradual transition from food collection to production is typical in most parts of the world.
Other Sites
Çatalhöyük
Asikli
Huts
Jericho, ’Ain Ghazal
Timber and reed huts
’Ain Mallaha, Jericho, Hallam Çemi
Figure 6.15 An artist’s reconstruction of houses from the early Neolithic in Southwest Asia. These houses had rock wall foundations with walls and roofs of timber and reeds covered with mud. In areas lacking rock, such as Abu Hureyra, the foundations were made of mud brick.
_ Site Abu Hureyra
217
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 218
Concept
<
Archaeobotany The study of prehistoric plant remains Preserved plants in archaeological sites are rare unless the remains have been carbonized, generally through burning or oxidation. Such burned plant materials can sometimes be obtained through a process called flotation. Excavated sediments are poured into a container of water (Figure 6.16), and the lighter, carbonized plant remains float to the top (Figures 6.17 and 6.18). In addition to the kinds of plants used, the major issues in archaeobotany (the study of the prehistoric use of plants) include the contribution of plants to the diet, medicinal uses, and domestication—the origins of agriculture. The archaeobotany of Southwest Asia is of particular interest because of the evidence for early domestication in this area (Table 6.2). Two varieties of wheat (emmer and einkorn), two-row barley, rye, oats, lentils, peas, chickpeas, and other plants were originally cultivated in Southwest Asia. The wild forms of these species are
TABLE 6.2 Common Food Plants in Early Neolithic Southwest Asia Einkorn wheat, wild and domesticated forms Emmer wheat, wild and domesticated forms Rye, wild and domesticated forms Barley, wild and domesticated forms Chickpeas, domesticated form Field peas, domesticated form Lentils, wild and domesticated forms Common vetch Bitter vetch Horse bean Grape, wild and domesticated forms Caper Prosopis (mesquite) Fig Hackberry Turpentine tree Wild pistachio
_ _
218
Chapter 6 The Origins of Agriculture
still common today, as they were in the past. Wild emmer wheat has a restricted distribution in the southern Levant. Wild einkorn wheat is relatively widespread in the northern and eastern sections of this region. Wild barley grows throughout the Fertile Crescent. All these wild grasses grow well in disturbed ground around human settlements. Einkorn was probably domesticated in southern Turkey, and emmer may have been first cultivated in the Jordan Valley. Agronomist Jack Harlan, of the University of Illinois, participating in an archaeological project in southern Turkey in the 1960s, experimented to find out just how much food was available from wild wheat. Dense stands of wild einkorn wheat grow on the slopes of the mountains in that area. This wild wheat is more nutritious than the hard winter red wheats grown in the United States today. Harvesting when the wheat was ripe, Harlan collected more than 1 kg (2 lb) of cereal grain per hour with his hands and even more with a sickle. He estimated that a family of four could harvest enough grain in 3 weeks to provide food for an entire year. If this wild wheat was so abundant and nutritious, why was wheat domesticated? The answer probably lies in the fact that wild wheats do not grow everywhere in Southwest Asia, so some communities may have transplanted the wild form into new environments. Although artifactual evidence for the use of plants (e.g., sickles, milling stones, storage pits, and roasting areas) exists in a number of areas, domesticated varieties cannot be distinguished from wild types without actual plant parts or grain impressions in clay bricks or pottery. Archaeobotanist Gordon
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 219
Hillman, of the University of London, studied wild einkorn and observed that simple harvesting had no major impact on the genetic structure of the wheat. Only when specifically selective harvesting and other cultivation techniques were applied could changes in the morphology of the seeds be noted. Such a pattern suggests that certain characteristics of domesticated wheat and barley, which show definite morphological differences from the wild ancestral forms, must have been intentionally selected. Results from Hillman’s experimental studies suggest that the change from wild to domesticated wheat may have occurred in a brief period, perhaps 200 years or less. According to Hans Helbaek, an archaeobotanist who worked on the issue of plant domestication, the most important characteristic of a domesticated species is the loss of natural seeding ability. The plant comes to depend on human intervention to reproduce. This change also permits humans to select the characteristics of those plants to be sown and reproduced, leading to preferred characteristics. Another major change in domesticated plants is the human removal of plants from their
Figure 6.16 A Dausman flotation machine in use. Water in the tank is used to separate lighter plant remains from soil and other sediments. The archaeologist uses a hose to spray some of the recovered materials.
Figure 6.17 In this photo, the charred seeds and charcoal separated by flotation are being captured in a cheesecloth sieve.
Figure 6.18 An assortment of seeds and other plant remains from an archaeological site.
flotation A technique for the recovery of plant remains from archaeological sites. Sediments or pit contents are poured into water or heavy liquid; the lighter, carbonized plant remains float to the top for recovery, while the heavier sediments and other materials fall to the bottom. archaeobotany (or paleoethnobotany) The study of plant remains
from archaeological sites.
_ Concept Archaeobotany
219
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 220
Wheat grain Glume Rachis
Figure 6.19 The important structural characteristics of wheat in the process of domestication.
rachis The stem that holds seeds to the stalk in wheat and other plants. glume The tough seed cover of many
cereal kernels. shattering A natural mechanism of
_ _
seed dispersal.
220
natural habitat and adaptation to new environmental zones. The distribution of the wild ancestor of einkorn wheat is shown in Figure 6.25. New conditions of growth would obviously select for different characteristics among the members of the plant species. Certain varieties do very well when moved to a new setting. Wheat is an annual grass with large seed grains that concentrate carbohydrates inside a hard shell. Grain at the top of the grass stalk is connected by the rachis, or stem. Each seed is covered by a husk, or glume (Figure 6.19). The major features that distinguish wild and domesticated wheat are found in the rachis and the glume. In wild wheat, each rachis of a seed cluster is brittle, to allow natural seed dispersal by a mechanism known as shattering. The glumes covering the seeds are tough, to protect the grain until the next growing season. These two features, however, are counterproductive to effective harvesting and consumption by humans. Because of
Chapter 6 The Origins of Agriculture
the brittleness of the rachis, many seeds fall to the ground before and during harvesting, making collection difficult. The tough glume must be roasted so that threshing can remove it. Domesticated wheats exhibit a reverse of those characteristics: a tough rachis and a brittle glume. These changes enable the seeds to stay on the plant so that they can be harvested in quantity and the glume can be removed by threshing without roasting. With these changes, the wheat is dependent on humans for seeding and therefore, by definition, is domesticated. One of the most exciting recent developments in paleoethnobotany has been the use of the scanning electron microscope. This instrument, with very high magnifications, has enabled researchers to identify minute scraps of charred plant remains that would otherwise be missed at archaeological sites. Electron microscopy also is being used to recognize edible plants such as roots and tubers, which were previously invisible in archaeological deposits.
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 221
Site
:
Jericho One of the oldest continuously inhabited places in the world
The walls of Jericho fell to Joshua and the Israelites sometime around 1300 B.C. The place had been occupied, however, for thousands of years before their arrival. In fact, the walls of Jericho tumbled down almost twenty times in the course of its history, from either earthquake or siege. Jericho is a tell, a massive mound of 2.4 ha (6 acres) composed of mud bricks and trash, accumulated to a height of 22 m (70 ft) during its long period of occupation (Figure 6.20). The lay-
ers of the tell built up at a rate of roughly 26 cm (almost 1 ft) every century. Jericho (or Tel es ‘Sultan, as it is known in Arabic) is one of the oldest continuously inhabited places on earth. Since at least 10,000 B.C., Elisha’s Fountain, the spring at Jericho, has witnessed virtually continuous human settlement. The freshwater spring floods the area beneath it and supports an oasis of luxuriant vegetation in the midst of the hot and arid Jordan Valley, lying
And it came to pass at the seventh time, when the priests blew with the trumpets, Joshua said unto the people, Shout; for the Lord hath given you the city . . . and the people shouted with a great shout. . . . The wall fell down flat . . . and they utterly destroyed all that was in the city . . . with the edge of the sword . . . and they burnt the city with fire, and all that was therein. —Joshua 6:16
Figure 6.20 The ancient tell of Jericho, located near a major spring at the northern end of the Jordan Valley in Southwest Asia. The tell is a mound of the accumulation of 10,000 years of human occupation.
_ Site Jericho
221
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Figure 6.21 A plastered and painted skull from Jericho. The plaster remodeling of the features of human heads is found at several early Neolithic sites in Southwest Asia and may reflect an increasing reverence for ancestors.
Page 222
275 m (900 ft) below sea level. The mound itself was abandoned sometime before A.D. 1, as settlement spread to the surrounding low area. The biblical connection has made Jericho a place of major interest and importance for archaeologists in Southwest Asia. Today, the tell resembles the surface of the moon. Craters and trenches mark the excavations of many archaeological projects that have explored these accumulated layers since 1873. British archaeologist Dame Kathleen Kenyon exposed a number of levels beneath biblical Jericho containing remains from the Bronze Age, the Neolithic, and the Mesolithic. Evidence from the early Neolithic (8500–7600 B.C.) at the bottom of the tell is of interest here. Because the mound is so deep, these lowest levels could be reached only by narrow trenches. The residential structures and artifacts exposed in Kenyon’s excavations were similar to those from other Southwest Asian sites from this period. Closely packed round houses contained interior hearths and grinding equipment.
50 c. 8000–7000 BC c. 7000–6000 BC
Percent of total diet
40
30
20
10
_ _
fox
pig
le catt
eep t /sh goa
s rum mall inan t
gaz
elle
0
Figure 6.22 Changes in the consumption of animals between 8000 and 6000 B.C. at Jericho. Notice especially the shift from wild gazelle to domesticated sheep and goat.
222
Chapter 6 The Origins of Agriculture
Headless burials were uncovered in the houses at Jericho; skulls were found separated from the skeletons (Figure 6.21). Kenyon estimated that the early Neolithic community at Jericho had a population of 600 people. Gazelle bones were abundant in the lowest layers at Jericho. This important game animal probably was exploited by large drives that captured a number of animals simultaneously. Aurochs, wild boars, and foxes also were eaten. Many animals present in the surrounding hills were not hunted, including wild goats, oryx, hartebeest, and wild camels. In later periods, gazelles declined in importance, replaced by sheep and goats (Figure 6.22). The presence of equipment for processing grain (sickles and grinding stones) and storage suggests that cereals were important in the lowest levels. In the slightly higher early Neolithic layers, these cereals were domesticated and probably cultivated in the fertile soil of the Jericho oasis. The inhabitants of early Neolithic Jericho traded various items over long distances. Materials such as salt, tar, and sulfur came from the area around the Dead Sea. Turquoise was brought from the Sinai Peninsula, cowrie shells from the Red Sea, obsidian from Turkey, and greenstone from Jordan. The reasons for such exchange probably lie in the importance of contacts with neighboring communities and in the accumulation of status items for some portion of the community. Long-distance exchange and expansion appear to have been hallmarks of the early Neolithic, probably reflecting the success of early agriculture and associated economic systems. Evidence for the spread of domesticates, architecture, and other components of the prepottery Neolithic seen at Jericho and elsewhere in the Levant appears on the island of Cyprus shortly before 8000 B.C. The colonists brought domestic cattle and wild deer with them to the island, traveling at least 80 km (50 mi) across open waters of the eastern Mediterranean. Most remarkable of Kenyon’s discoveries from this period was a large
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
stone tower (Figure 6.23), wall, and ditch, which appeared to encircle the site. These structures were built at the beginning of the eighth millennium B.C. The wall itself is 1.8 m (6 ft) thick at the base, narrowing to 1.1 m (3.5 ft) at the top, and stands 3.6 m (12 ft) high today, buried under the accumulated deposits of Jericho. The rubble-filled stone tower also was completely buried, at a height of 8.2 m (27 ft). The tower was built inside the wall and has a diameter of 9 m (30 ft) at the base. An interior staircase with 22 steps leads from the bottom to the top. In front of the wall is a deep ditch, which added greatly to the height of the total structure. The ditch is 2 m (6.5 ft) deep and 8.5 m (28 ft) wide, cut into the bedrock under the site. Such a construction project would have been a major undertaking by a small community in the eighth millennium B.C. The tower and wall were abandoned by 7300 B.C. and rapidly disappeared under the accumulating layers of the tell. The significance of these structures has been the focus of discussion and debate. Kenyon suggested that the tell was completely encircled by the wall and ditch, with towers placed at regular intervals. However, only one tower was exposed in the excavations, and the wall was not found on the southern, lower, side of the mound. Numerous unanswered questions have arisen regarding the nature of the enemy, the materials to be protected, the location of the tower inside the wall, the reason for its abandonment, and the size of the population. Comparable fortifications are not seen at other early Neolithic sites in Southwest Asia, and a massive defensive structure seems out of place. Ofer Bar-Yosef, of Harvard University, examined the evidence for fortifications at Jericho and found it lacking. Bar-Yosef interprets the wall and ditch as a defense against nature rather than against humans (Figure 6.24). In this arid region, flash floods are common, and each rainfall moves a great deal of sand and silt from higher ground to lower areas. Such erosion and deposition would have been aggravated by
Page 223
land clearing and the removal of trees and shrubs by the inhabitants at the site. Sediments may have accumulated on the upslope side of the tell of Jericho and at other early Neolithic sites. To counteract this rapid accumulation of sediments at the edge of their community, the residents of Jericho, Bar-Yosef argues, built the wall and dug the ditch to hold back the sediments. The rock-cut ditch at Jericho filled with water-transported sediments shortly after it was dug. The wall was thickest in the middle portion, where the sediment accumulation was heaviest. The large tower inside the wall is another matter. Bar-Yosef has noted the excellent preservation of the tower to its total height in the tell. He argues that there was likely a mud-brick structure on top of the stone tower, a building for communal storage or religious activities that made the tower a very special place in the settlement. The tower thus may represent an early shrine or temple for the community, a concept that was later transformed into the monumental ziggurats of the Mesopotamian states. Around 7500 B.C., major changes in the architecture, artifacts, and animals occurred at Jericho. Houses became square, constructed with what have been described as “hog-backed” bricks—large, cigar-shaped mud bricks with a herringbone pattern of thumbprints on the top. Domesticated animals became important at that time; sheep and goats constituted almost half of the animal bones, along with a few wild cattle and pigs, gradually replacing the gazelle. Pigs and cattle were possibly in the process of change, because their bones were slightly smaller than those of the wild forms but not as small as those of the domesticated forms.
Figure 6.23 The tell at Jericho. Excavations have exposed a circular stone tower.
Figure 6.24 The tower, wall, and rock-cut ditch at Jericho. Internal stairs Tower
Entrance
Later wall
Ditch Bedrock
Site Jericho
_
223
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 224
Concept
<
Archaeozoology The study of animal remains
y
Why were plants and animals domesticated about the same time?
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on archaeozoology, see the Internet exercises on your online learning center.
The companion of paleoethnobotany, archaeozoology focuses on the hard body parts of animals that survive—bone, teeth, antler, ivory, scales, and shell. Studying these materials, archaeozoologists attempt to answer questions about whether animals were hunted or scavenged, how animals were butchered, how much meat contributed to the diet, and the process of domestication. Archaeozoologists are trained to identify the genus or species of an animal from small fragments of bone, as well as the age and sex of the animals, how the animal was butchered and the bone was broken, and how many individual animals are represented in the bone assemblage. Fracture patterns in long bones may reveal intentional breakage for removing marrow. An analysis of cutmarks on bone may provide information on butchering techniques. The study of animal domestication is also an important part of archaeozoology. Four major criteria are used to look
Figure 6.25 The distribution of wild einkorn wheat (Triticum boeoticum) and of wild goats (Capra aegagrus) in Southwest Asia and Egypt is indicated by the shaded areas.
for domesticates: geographic evidence, abundance, morphological changes, and herd demographics. Geographic evidence involves discovery of animal species outside their natural habitat and presumes human involvement. However, environments changed dramatically in the past, as did the geographic distribution of animal species. Thus geography is a difficult criterion to use. Increases in the abundance of a species in the layers at a site are often taken to indicate domestication, but again this evidence is not particularly reliable. Numbers of animals may increase for a variety of reasons, including environmental change and increased hunting. Herded animals show certain morphological changes in size and body parts that provide direct evidence for domestication. Domesticated species are generally smaller than their wild ancestors. The shape of the horns often changes in the domestic form, and the microscopic structure of bone under-
ARMENIA
TURKEY
AZERBAIJAN
Hallam Cemi Catalhöyük Abu Hureyra
ISRAEL
SYRIA
Jericho
IRAN
IRAQ
JORDAN EGYPT
SAUDI ARABIA
KUWAIT
Distribution of wild goats (Capra aegagrus) archaeozoology The study of animal remains from archaeological sites.
224
Caspian Sea
CYPRUS Mediterranean LEBANON Sea
_ _
KAZAKHSTAN
0
Chapter 6 The Origins of Agriculture
300 km
Distribution of wild einkorn wheat (Triticum boeoticum)
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 225
Catastrophic (living structure) mortality profile
Attritional (U-shaped) mortality profile
Prime-dominated mortality profile
Number of individual animals
10
8
6
4
2
0
1
2
3
4
5
6
7
8
9
Age at death (years)
goes modification in domestic animals. Other traits may be selected by herders to increase the yield of milk, wool, or meat. However, because such biological change takes many generations and requires physical separation between wild and domestic animals, the earliest stages of domestication may not have been recorded in the bones and horns that remain (Table 6.3). Brian Hesse, of the University of Alabama–Birmingham, and Melinda Zeder, of the Smithsonian Institution, have used herd demographics to document early animal domestication in Southwest Asia (Figure 6.26). They estimated the age and the sex of animals that had been killed at prefarming sites in the Zagros Mountains of western Iran and used this information to study whether hunting or herding was practiced. The basic principle relies on the fact that herded animals are slaughtered when the herder decides; for most species, this means that the average age of death for domesticated animals is younger than for wild animals. Hunted animals are killed in chance encounters, and the proportion of adults is higher in such situations. The ages of animals are most frequently determined by an assessment of tooth eruption and wear, along with information about changes in bone. All
1
2
3
4
5
6
7
8
9
Age at death (years)
the known sites in the Zagros Mountains before 10,000 years ago show similar slaughter patterns for sheep, goats, and red deer; bone assemblages contain primarily adult animals, indicating that all were hunted in the wild. However, a number of sites contain assemblages after 10,000 years ago that are dominated by the bones of younger animals. In each case, the younger groups are sheep or goats, proportionally higher than in a normal wild herd. At Abu Hureyra, a study of the bones of wild gazelles and domesticated goats and sheep has provided new information on the process of animal domestication (Figure 6.27). Around 11,000 years ago, the site was occupied by prefarming huntergatherers who hunted gazelles in large numbers. The gazelle bones and teeth include the remains of many young animals. The teeth, in particular, indicate that both newborns and yearlings were common in the faunal assemblage, along with adults of all ages. This pattern of newborns, yearlings, and adults, and the absence of animals of ages in between, indicates that most of the animals were killed during the same time each year, shortly after the calving period in late April and early May. These hunters were taking entire herds of gazelles as the animals migrated north
1
2
3
4
5
6
7
8
9
Age at death (years)
Figure 6.26 Three hypothetical age profiles for animal populations used by archaeozoologists to look for indications of human control and domestication. A catastrophic pattern would reflect that all the animals in the herd died at the same time. An attritional pattern is the normal lifeand-death cycle for an animal herd in which very young and very old animals are more likely to die. The prime-dominated mortality curve is the pattern that appears when humans eat a predominance of young males, a common herding practice. Females and a few older males are kept for the reproduction of the herd.
_ Concept Archaeozoology
225
10/1/09
8:02 AM
Figure 6.27 Changes in animal species at Abu Hureyra, Syria, between 9000 and 6000 B.C. The width of the bars indicates the relative abundance of the species. The most pronounced change is the decrease in gazelles and the increase in sheep and goats between 6500 and 6000 B.C. There was an absence of human occupation at this site between 8000 and 7500 B.C.
Page 226
Gazelle
Sheep and goat
6000
Wild ass
Wild cattle
Wild pig and deer
6500
7000 Years B.C.
pri31057_ch06.pp198-265.qxd
7500 Break in human occupation 8000
8500
9000
0
20
40
60
80 0
20
40
60
80 0
20 0
20 0 10
Percentage of total bones
during early summer. They probably used a technique that would drive an entire herd into an enclosure or series of pitfalls, where all the animals could be killed. These hunters were so effective that the number of gazelles in the area dropped dramatically, to less than 20% of all the animals at the site, by 7500 B.C. Goat and sheep domestication may have been a solution to the problem of decreasing numbers of gazelles. Sheep and goats were slaughtered throughout the year, in contrast to the seasonal hunting of gazelles at Abu Hureyra. Goat and sheep bones constituted about 10% of the faunal assemblages until 7500 B.C., after which they very rapidly became the predominant component, at almost 80%. This period was about 1500 years after plant domestication had been initiated at the site; wheats and barley at that point provided a significant portion of the diet. It also was some 3500 years after the initial occupation of the site, documenting the sequential stages of sedentism, plant cultivation, and animal domestication typical of Southwest
_ _
226
Chapter 6 The Origins of Agriculture
Asia. Such information suggests a very complex picture of animal and plant domestication in Southwest Asia at the end of the Pleistocene.
TABLE 6.3 Important Animal Species in Neolithic Southwest Asia Gazelle Goat, wild and domesticated forms Sheep, wild and domesticated forms Roe deer Fallow deer Cattle, wild and domesticated forms Pig, wild and domesticated forms Onager Bear Jackal Hare Wildcat
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 227
Site
:
Çatalhöyük The first city, central Turkey
Large communities began to appear shortly after the domestication of plants and animals in Southwest Asia. By 8000 B.C., ’Ain Ghazal and Jericho had populations in the hundreds, sizably larger than those in pre-agricultural settlements. And by 7250 B.C. the first “city” had appeared at the site of Çatalhöyük (sha-TAL-who-YUK) in central Turkey (Figure 6.28). The tell of Çatalhöyük is huge, 600 m (1900 ft) long, 350 m (1000 feet) wide, and almost 20 m (65 ft) high (Figure 6.29). This massive mound of houses, garbage, and burials accumulated within a period of little more than 1000 years and was abandoned around 6000 B.C. At least twice as large as early Neolithic Jericho, covering 13 ha (32 acres), Çatalhöyük was a large settlement of perhaps as many as 2000 families—on the order of 10,000 people. The original excavations at this ancient settlement were conducted by James Mellaart, of the University of London, during the 1960s. Several
Figure 6.28 Stratigraphy (top) and radiocarbon dates (bottom) from a deepsounding at Çatalhöyük. The section drawing shows 4 m of the complex stratigraphy and the location of the radiocarbon samples. The graph of AMS radiocarbon dates shows the probability curves for the age of each sample. The center of each line of hills is a good estimate for the age of the sample. The dates indicate that the oldest occupation at Çatalhöyük was around 7250 B.C.
East
West A and B C
D and E
F G
H
I
J
K L
0
1m
A
7935±50BP
B
7980±55BP
C
7985±50BP
D
8090±55BP
E
8090±50BP
F
8050±50BP
G
8030±50BP
H
7985±55BP
I
8150±50BP
J
8160±50BP
K
8240±55BP
L
8155±50BP
8000Cal BC
7500Cal BC
7000Cal BC
6500Cal BC
Calibrated date
_ Site Çatalhöyük
227
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 228
Figure 6.29 An aerial view of the Neolithic mound of Çatalhöyük. Paths on top of the mound lead from the excavation headquarters to major excavation areas.
y
What role does sedentism play in the transition to agriculture?
_ _
228
seasons of fieldwork at the site exposed numerous walls and floors of rooms and houses. Houses were built closely together in one, two, or three stories around small courtyards (Figure 6.30). The houses were very similar, with a rectangular floor plan of approximately 25 sq m (30 sq yd), about the size of a large living room today. The houses were divided into a living area and a smaller storage area. Furniture —benches, sleeping platforms, ovens, cupboards, and storage bins—was built into the house. The houses had no doors, and access was likely through their flat roofs. A number of burials also were found in the houses at Çatalhöyük. These burials of men, women, and children were under the floors and sleeping platforms (Figure 6.31). The bodies appear to have been exposed for some time before burial. Grave goods with the burials included jewelry such as necklaces, armlets, wristlets, and anklets of stone or shell, copper and lead beads, and weapons. A few burials contained rare objects such as stone vessels, ceremonial daggers, obsidian mirrors, polished maceheads, cosmetics, and metal beads and rings.
Chapter 6 The Origins of Agriculture
One of the remarkable things from the first excavations at Çatalhöyük was the large number of the structures, perhaps 20%, that appeared to have been shrines (Figure 6.32). The walls of these shrines are elaborately decorated, sculpted, and painted with a variety of remarkable figures and designs, including vultures, bulls, wild cats, and humans. Some of the paintings show women giving birth to bulls; others depict hunting scenes or vultures with headless humans in their talons. One of the paintings portrays an erupting volcano with a large settlement at its base. In addition, a number of sculptures and figurines have been excavated (Figure 6.33). New excavations in the 1990s exposed more of the site and investigated in more detail the purpose and function of the various structures. These excavations, directed by Ian Hodder of Stanford University, suggest that households used their space for both domestic and ritual purposes and that the shrines may simply have been more elaborate households. Two or three generations of a family were often buried under the house floor. The first burials in the houses were infants and young children; later
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 229
Figure 6.30 An artist’s reconstruction of complex architecture at Çatalhöyük, showing the closely packed, multistory buildings of timber and mud brick. This view reconstructs only a small part of the settlement.
burials were older adults. This pattern suggests a family life cycle represented in the burials. Houses may have been destroyed after the family had died. The paintings and sculptures on the walls of the houses are likely associated with the burials and may have been added to commemorate the deceased. David Lewis-Williams, of the University of Witwatersrand, South Africa, has described the artwork as a symbolic membrane connecting the living to the spirit world. Analysis of the animal bones and plant remains has provided much new information about the site. The inhabitants depended heavily on wild flora and fauna. Important plants in the diet included domesticated wheat and barley, wild tubers and grasses, lentils, and fruits and nuts such as acorns, pistachios, crab apples, and hackberries. Cattle were an important part of subsistence at the site, but it is not yet certain if they were domesticated. Domesticated sheep and other species also were killed and eaten. There are no indications of differences in status, represented by wealth or surplus, among the houses. Çatalhöyük appears in
many ways to have been a huge village of farmers rather than a complex and varied city. Çatalhöyük was clearly a prosperous center, however, probably because of its control of the obsidian trade. Obsidian is produced by volcanoes; molten silica sometimes flows out of a volcanic core and hardens into this stone, which was highly sought by prehistoric makers of stone tools. Obsidian, like glass and flint, fractures easily and regularly, creating very sharp edges. In the past, obsidian was often traded or exchanged over long distances—hundreds of kilometers or more. It is available from only a few places, limited by proximity to volcanic mountains and the chance formation of a silica flow. Most sources for obsidian are known because they are rare and the material is unusual. It also is possible to fingerprint different flows of obsidian through minor differences in the chemical composition of the material, which is specific to each source, allowing pieces found elsewhere to be traced to the places where they originated.
Agriculture quickly changes the way human society looks archaeologically. What are some of those changes?
y
obsidian Translucent, gray-to-black or
green, glasslike rock from molten sand.
Site Çatalhöyük
229
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 230
Figure 6.31 A group of burials beneath a platform in one of the houses.
Figure 6.32 An artist’s reconstruction of one of the shrine rooms at Çatalhöyük. A catlike goddess gives birth to a ram above three bulls’ heads.
_ _
230
Chapter 6 The Origins of Agriculture
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
The sources of obsidian in Southwest Asia, the Aegean area, North America, Mexico, and elsewhere have been studied using such methods. Most of the obsidian in Southwest Asia comes from sources either in the mountains of Turkey or in northern Iran, both outside the Fertile Crescent. The sources of obsidian found at early Neolithic sites provide information on both the direction and the intensity of trade. Sites in the Levant generally obtained obsidian from Anatolia; sites in the Zagros used Armenian material. The percentage of obsidian in the total flaked stone assemblage at these sites indicates that sites closest to the sources used a great deal of obsidian, whereas those farthest away obtained only a small amount. At Jericho, for example, 700 km (400 mi) from the Turkish sources, only about 1% of the stone tools were made from obsidian. Çatalhöyük is located almost 200 km (125 mi) from the major obsidian source in Turkey. Nevertheless, most of the chipped stone tools at the site were made of obsidian. In addition to finished tools, many unfinished obsidian artifacts were found, along with large amounts of raw material. It appears that obsidian was moved in huge quantities to Çatalhöyük. From here the obsidian was traded over a wide area of Southwest Asia. In return, the inhabitants of Çatalhöyük received copper, shell, and other exotic materials. Clearly, the importance of obsidian as a desired object in trade was an essential factor in the rise of Çatalhöyük. Trade and exchange of various materials accelerated greatly during the Neolithic. Remarkable as Catalhöyük is, there is another site, 100 km (62.5 mi) distant, that is even more remarkable. Although this mound, called Asikli, is 1000 years older than Çatalhöyük, it is
Page 231
Figure 6.33 Two clay female figurines from the excavations at Çatalhöyük (front and side view of each).
almost as tall. This site contains little evidence for domesticated plants or animals. At Asikli, a population of several hundred individuals lived in a large group of mud-brick houses surrounded by a stone wall. There are at least ten levels of occupation in the mound, and the same arrangement of houses and cobbled streets is seen in each level. There is also a cluster of larger, public buildings at the site that might have been a temple complex.
_ Site Çatalhöyük
231
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 232
Site
:
Mehrgarh New evidence for the early Neolithic in South Asia
Figure 6.34 Multiroom, rectangular structures uncovered during excavations at Mehrgarh.
_ _
Figure 6.35 A grave from an early level at Mehrgarh containing an adult male buried with several body ornaments and five young goats.
232
The Indus civilization, one of the world’s early urban societies, emerged in South Asia, the subcontinent now composed of India, Pakistan, Bangladesh, and Nepal. The largest communities of this ancient civilization—the cities of Harappa, Mohenjo-daro, and others— were centered on the Indus River system in Pakistan. Until the 1970s, however, little was known about the antecedents of this ancient society. In the late 1960s, it was suggested that plant and animal domestication did not reach the subcontinent of South Asia until after 4000 B . C . The prevailing opinion was that migrants from the west, who made metals and wheelthrown pottery, brought an agricultural way of life to South Asia just a few centuries before the rise of the Indus civilization. More recent archaeological findings, however, have revealed an older, more indigenous picture of agricultural origins in the Indus River drainage. One key for this new perspective has come from a long-term research program, conducted by the French Archaeological
Chapter 6 The Origins of Agriculture
Mission and the Pakistani Department of Archaeology, at the site of Mehrgarh (meh-her-GAR), located in the Kachi Plain about 200 km (120 mi) northwest of the Indus River. The site is also interesting because of its location immediately below the Bolan Pass, which cuts through the mountains that connect the Indus River Valley with highland Baluchistan and Iran (Southwest Asia). In 1974, the first fieldwork at Mehrgarh, by Jean-François Jarrige and his colleagues, focused on a small mound. At that time, however, older pottery was collected over an area of several hundred hectares adjacent to the mound. Largescale excavations in this area have yielded a sequence of deposits dating to the seventh millennium B.C. The earliest occupation level at Mehrgarh lacked pottery, although clay was used to make bricks for the construction of substantial, multiroom, quadrangular structures (Figure 6.34) and anthropomorphic figurines. A count of plant impressions in the mud bricks by Lorenzo Costantini found that barley was the most abundant cultigen in this early level. This barley had several distinctive local characteristics and may not have been completely domesticated. Other cereals grown at that time include smaller amounts of einkorn wheat, emmer wheat, and a kind of bread wheat. Whereas the wheats appear to have been imports from the west, it remains unclear whether barley was as well. As in the early Neolithic sites of Southwest Asia, gazelle was the most abundant of the wild animal species, which included sheep, goat, cattle, swamp deer, and a large South Asian antelope. Some goats have been identified anatomically as domesticated. In each of two burials from an early level of the occupation, five young goats were placed at the foot of an adult male, whose body was covered with red
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
ochre (Figure 6.35). These early graves also contained a diverse combination of body ornaments made from exotic materials such as seashell, turquoise, and lapis lazuli. After 6000 B.C., important subsistence shifts occurred at Mehrgarh, and the first ceramic vessels began to be used. Richard Meadow, of Harvard University, has noted a reduction in the body size of the goats, sheep, and humped zebu cattle (Bos indicus) at the site, which he interprets as evidence for domestication. The relative abundance of wild species decreased, and the proportion of cattle bones increased greatly, suggesting that cattle husbandry may have begun about the same time at Mehrgarh as in sites to the west (Figure 6.36). Therefore, whereas sheep and goats may have been brought to Mehrgarh already tamed or domesticated, archaeological as well as recent genetic evidence suggests that local humped cattle were indigenous to South Asia. Between 6000 and 4000 B.C., domesticated barley, well adapted to floodplain irrigation, was the predominant cultigen. Charred seeds of the plumlike jujube fruit (Zizyphus jujuba) and date pits (Phoenix dactylifera) also have been recovered. Cotton seeds were found with wheat and barley grains after 5000 B.C., the earliest date for cotton in the world. At that time, Mehrgarh was a wellplanned community composed of compartmentalized, mud-brick structures that served primarily as storage rooms. Features found in other parts of the settlement include circular fireplaces, containing burnt rocks for stone boiling, with bone and other debris nearby. These areas may have been used for large-scale food processing or cooking, or some other kind of communal activity. Both specialized craft production and extensive long-distance trade, so evident at the major centers of the later Indus civilization, had clear antecedents millennia earlier at Mehrgarh. By 4000 B.C., the community had spread over tens of hectares and included specialized centers where fine ceramic ware was made. For roughly the next 1500 years, Mehrgarh was an important craft center. Wheel-thrown pottery
Page 233
Other wild Wild Domesticated
Swamp deer Antelope Upper deposits
Gazelle Cattle Sheep Goat Other wild Swamp deer Antelope
Middle deposits
Gazelle Cattle Sheep Goat Other wild Swamp deer Antelope
Lower deposits
Gazelle Cattle Sheep Goat 0
10
was mass-produced, and beads of lapis lazuli, turquoise, and carnelian were perforated with cylindrical drill bits made of jasper and rotated by a bowdrill. Fragments of crucibles used to melt copper also were found at Mehrgarh. The archaeology at Mehrgarh, along with recent studies at other contemporary sites, has revised South Asian prehistory. No longer can we envision the inhabitants of the Indus River drainage as simple recipients of inventions from the west. Nor can we attribute the rise of the Indus civilization after 2600 B.C. to the diffusion of ideas from Mesopotamia. We now recognize a sequence in South Asia that documents a prepottery Neolithic phase and the indigenous domestication of humped zebu cattle, along with the development of highly specialized local craft industries.
20
30 40 Percentage of total
50
60
Figure 6.36 The transition from hunting to herding at Mehrgarh. In the lower deposits, all the species present are wild, including gazelle (Gazella), goat (Capra), sheep (Ovis), cattle (Bos), antelope (Boselaphus), and swamp deer (Cervus). After 6000 B.C. (middle deposits), the reduced size of cattle, goat, and sheep bones at Mehrgarh indicates that these animals were domesticated. After the Neolithic (upper deposits), domesticated cattle were dominant.
anthropomorphic Having human
form or attributes. lapis lazuli A semiprecious stone of
deep blue color. stone boiling The process of heating
stones in a fire and then adding them to containers to boil water or cook other foods. jasper A high-quality flint, often
highly colored. bow-drill A device for perforating
beads or other small objects, in which a bow is used to rotate the shaft of the bit.
Site Mehrgarh
233
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 234
Concept
<
Pottery Ancient containers: a key source for archaeological interpretation
Figure 6.37 A Mexican woman making a pottery vessel using a molde, a flat, roughly finished clay plate that has traditionally been used in parts of Mexico to turn vessels by hand.
“Pottery is . . . the greatest resource of the archaeologist,” wrote the famed Egyptologist W. M. Flinders Petrie (1904, pp. 15–16). Indeed, ceramics are the most common kind of artifact found at most post-Paleolithic sites. Since pottery has many purposes— cooking, storage, serving, and carrying materials—many different pieces can be used by a single household at the same time. Moreover, pottery vessels are fragile and often have to be replaced. However, fragments of pottery, or potsherds, are very durable and normally preserve better than many other ancient materials found in archaeological contexts.
_ _
234
Chapter 6 The Origins of Agriculture
Ceramic artifacts also are important because they can be good temporal markers, sensitive indicators of specific time periods. In his study of ancient Egyptian pottery, Flinders Petrie was one of the first archaeologists to recognize how decorative styles change. In addition to chronological sensitivity, pottery vessels have a series of distinctive technical, formal, and decorative attributes that can tell us many things about the lives of the people who made, traded, and used them. Worldwide, the increasing importance of pottery has in many cases roughly coincided with a greater reliance on domesticated foods. Fired clay containers provide clean storage for food and drink and can be used for preparing food over a fire. The earliest securely dated pottery vessels, found at Jomon sites in Japan, are 10,000–12,000 years old (see “Sannai Maruyama,” Chapter 5, p. 183). There may be pottery of roughly the same age from several areas of mainland China as well. Why did they occur so late in human history? Ceramic vessels do have liabilities: They are relatively heavy and often fragile. Many mobile foragers did not use ceramics, preferring lighter containers, such as net bags, gourds, and baskets. In the preceramic levels at Mehrgarh, for example, the inhabitants used baskets coated with water-resistant bitumen. Clearly, the development of and reliance on ceramic vessels is associated with life in more settled communities. Yet some sedentary groups, such as the Native Americans of northwestern North America, did not use pottery, relying instead on a diverse array of woven baskets for storage. And pottery did not simply appear with agriculture. In both Mexico and Southwest Asia, plant domestication preceded the use of ceramics by more than a thousand years. Conversely, the
pri31057_ch06.pp198-265.qxd
10/15/09
10:53 AM
Page 235
Figure 6.38 Early pottery forms in Mexico were based on cross sections of gourds.
world’s oldest ceramic containers were made by the nonagricultural Jomon fisher-foragers of ancient Japan. The late advent of pottery is curious because ceramic technology had been used by human societies for some time. Baked clay figurines were made at Paleolithic sites in the former Soviet Union and eastern Europe as early as 30,000–20,000 years ago. Clay figurines and mud bricks were made in the Levant 10,000 years ago, yet pottery did not appear locally until around 7000 B.C. Settled farming villagers used ground stone bowls long before they made ceramic containers. In the Americas, ceramic vessels were not made until several thousand years later, when they first appeared in lowland South America. The earliest pottery containers in Mexico and North America date to 3000–2000 B.C., before the advent of fully sedentary villages. The earliest pottery in both Eurasia and the Americas was crafted by hand (Figure 6.37). Generally, early pottery vessels were either built up in a series of clay coils or modeled from gourds (Figure 6.38). However, by roughly 4000 B.C., wheel-thrown pottery had become an important commodity throughout Southwest and South Asia. The potter’s wheel permits a single worker to make a greater number of vessels more quickly. Such pottery is also highly uniform in size, shape, and appearance. In the Americas, fired clay vessels continued to be handmade (sometimes including the use of molds; Figure 6.39) until the European introduction of the potter’s wheel in the sixteenth century A.D. Some scholars have
Figure 6.39 A funerary urn from a prehispanic tomb in Oaxaca, Mexico, dating to A.D. 600–800. This elaborate urn stands 15 cm (6 in) high and was made using a variety of small molds to form the arms, the headdress, and other elements that were then attached to the main vessel.
attributed the Eurasian invention of the potter’s wheel to the presence of wheeled vehicles, absent in the Americas. Yet the potter’s wheel may have preceded the use of the wheel for transport in Eurasia. Furthermore, we know that some Native American groups were familiar with the concept of the wheel, because wheeled toys have been found in archaeological contexts. The absence of the potter’s wheel may be related to the relative difficulty and inefficiency of transporting ceramic vessels long distances in the aboriginal Americas, where there were few domesticated beasts of burden, no wheeled vehicles, and generally less widespread sea transport than in Eurasia and Africa.
Compare and contrast the context in which fired clay vessels were adopted or developed in three areas of the world.
y
potsherd A fragment of a clay vessel
or object. temporal marker A morphological type, such as a design motif on pottery or a particular type of stone tool, that has been shown to have a discrete and definable temporal range. wheel-thrown pottery Pottery that is made using the potter’s wheel.
Concept Pottery
235
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 236
Site
:
Ban-po-ts’un A Neolithic village in northern China Botanists have long recognized that many Chinese food plants were indigenous to the region. Prior to the last half century, however, archaeologists believed that the “idea” of agriculture, along with the knowledge for domesticating animals and making pottery, had diffused to the rich soil of the middle Huang (Yellow) River Valley in northern China from elsewhere. Before 1960, the few known Neolithic sites with
Figure 6.40 Location of Ban-pots´un and other sites mentioned in the text. The distribution of Yangshao sites is shown in yellow.
Sea of Japan JAPAN MONGOLIA
KOREA
Huang River
Chiang-chai YANGSHAO Ban-po-ts’un
e
Ho-mu-tu Shang-shan
Peng-tou-shan
Ya n
g
tz
v er Ri
PACIFIC
CHINA
MYANMAR (BURMA)
OCEAN
LAOS South China THAILAND VIETNAM
_ _
Sea
CAMBODIA
236
Chapter 6 The Origins of Agriculture
PHILIPPINES
their characteristic black-and-red pottery were assigned to the Yangshao culture and presumed to date to around 3000 B.C. The general view that products from Southwest Asia arrived in China about that time was supported by the presence of small amounts of foreign domesticates, such as wheat and barley, at known Yangshao sites, along with larger quantities of a locally domesticated grain (millet) and domesticated pigs. In the past 40 years, however, information about the Chinese Neolithic has undergone a major transformation. Thousands of new Neolithic sites have recently been discovered in various areas of the country. Dozens of these sites have been excavated, and several earlier sites have been recorded (Figure 6.40). Early Neolithic sites in North China with millet and pigs have now been dated to the sixth millennium B.C., and the age for some Yangshao sites has been pushed back to 5000 B.C. Equally significant has been the discovery of Neolithic sites, such as Peng-tou-shan, Shang-shan, and Ho-mu-tu in South China, that predate 6000–5000 B.C. At these more southerly sites, the staple food plant was rice, not millet. Charred husks of domesticated rice from Shangshan, a small sedentary village in the lower Yangtze River valley, have been dated to 8000 B.C. The origin of agriculture now appears to have been an indigenous process in North and South China, since in both areas local cultigens have been found to be more abundant and earlier than exotic domesticates. The best-known Chinese Neolithic site, Ban-po-ts’un (ban-POT-sun), is not the oldest but rather the first to have been excavated extensively. (Work was completed between 1953 and 1955.) Located on a loess terrace about 9 m (30 ft) above a tributary of the Huang River near the city of Xian, Ban-po-ts’un cov-
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
ers 5–7 ha (12.5–17.5 acres). Roughly 100 houses, some circular and others square, were surrounded by a defensive and drainage ditch (Figure 6.41). Many of these structures were excavated, and the evidence indicates that the occupation at Ban-po was long and continuous. In one instance, five superimposed house floors were uncovered. Many of the houses were semisubterranean, 3–5 m (10–16 ft) in diameter, with floors roughly a meter below the ground surface. Each house had timber beams that rested on stone bases, supporting a steeply pitched thatched roof. Floors and interior walls were plastered with clay and straw. One or two circular or pear-shaped fire pits, modeled in clay, were situated at the center of most of the dwellings. Storage pits and animal pens were interspersed among the houses at the center of the settlement. At Ban-po, the principal crop was millet (Setaria italica), which was cultivated in the rich, soft-textured loess soils that surround the village. Such agricultural tools as bone hoes, polished stone adzes, axes, knives, and digging-stick weights were abundant at the site. Chestnuts, hazelnuts, and pine nuts supplemented the grain diet. The inhabitants of Ban-po-ts’un grew hemp, probably for use as a fiber. Silk production is suggested by a neatly sliced silkworm cocoon that was recovered. Numerous spindle whorls, for spinning thread, and bone needles also were found. Impressions of cloth, as well as baskets and mats, provide more evidence of weaving. Pigs and dogs were the principal domesticated animals, although cattle, sheep, and goats also were utilized. Bone and quartz arrow points, bone fishhooks, and net-sinkers all were present, along with plentiful bones of several varieties of deer. Thus hunting and fishing contributed to the diet at Ban-po. Ban-po-ts’un has yielded more than 500,000 pieces of pottery. Six pottery kilns were recovered beyond the ditch at the east side of the settlement, outside the residential zone (Figure 6.42). Most of the vessels were handmade into a distinctive redware.
Page 237
Whereas cooking pots tended to be coarse and gritty, water vessels and food-serving bowls were made of a finer clay. Cord-marking was the most common surface decoration, although basket, textile, and fingernail impressions also were used. The black-painted geomorphic and zoomorphic Yangshao designs were applied primarily to bowls and jars (Figure 6.43). The inhabitants of Ban-po were buried in one of two ways. Infants and small children were placed in large redware pottery jars and interred near the houses. The cemetery for adults was located outside the enclosing ditch at the north end of the settlement, where corpses were placed in pits 2 m (6.5 ft) deep and arranged in rows. With very few exceptions, each individual was buried separately in an extended position. Ceramic vessels
Figure 6.41 An artist’s reconstruction of houses at Ban-po-ts’un. Circular structures are believed to have housed a single family; the larger, square structure is thought to have been a communal clan house.
hemp A tall annual plant whose tough fibers are used to make coarse fabrics and ropes. spindle whorl A cam or balance
wheel on a shaft or spindle for spinning yarn or thread from wool, cotton, or other material; usually made of clay. kiln A furnace or oven for baking or drying objects, especially for firing pottery. geomorphic Having the form or at-
tributes of surface features of the earth or other celestial bodies. zoomorphic Having animal form or
attributes.
Site Ban-po-ts’un
_
237
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 238
Figure 6.42 A drawing of a reconstructed section of a pottery kiln at Ban-po-ts’un.
were included with the body in most of the graves. In a few instances, larger quantities of goods were found. The most elaborate burial was a child, who was placed in a wooden tomb that included a green jade pendant, a string of 63 bone disk beads, four ceramic vessels, and three stone pellets. Toward the end of the occupation at Ban-po, a large rectangular structure was erected on a manmade platform (20 × 12.5 m, or 65 × 41 ft) at the center of the village. The platform was ringed by a low wall that originally may have been the foundation for a higher wall of posts. Unlike the residential dwellings, this structure was plastered with a white limy substance that had been hardened by baking. The structure also had a hard earthen floor that appears to have been destroyed in a fire. Although some archaeologists might consider this special, central structure to be a possible indicator of emergent social inequalities, Chinese archaeologists interpret the building as a communal assembly hall or clan house. The purposefully planned layout of Ban-po-ts’un, with a large central building surrounded by a nucleated residential area inside a ditch, and exterior pottery-making and cemetery facilities, was very similar to other contemporary Yangshao villages. For example, Chiang-chai, near Ban-po, had concen-
_ _
238
Chapter 6 The Origins of Agriculture
Figure 6.43 Examples of black-painted zoomorphic and geomorphic designs on Yangshao pottery.
tric circles of oblong and rectangular houses surrounding four large structures that faced a central village plaza. Chiang-chai also was ringed by a defensive system of ditches, and the cemetery area and pottery kilns were located outside these earthworks. At all wellstudied Yangshao sites, the dwelling and cemetery areas are spatially distinct.
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 239
Concept
<
Rice One of the world’s most important grains
Today, the most important staple food for half of the world’s population is rice. Each year, the plant supplies 20% of the total calories consumed worldwide. In fact, among the world’s major grain crops, rice is the only plant that is harvested almost solely for human consumption. Yet despite its worldwide significance, archaeologists know relatively little about how or where rice was first domesticated. Although the plant is now cultivated in a variety of environments, it was originally indigenous to the tropics. Rice is an annual grass that shares many characteristics with wheat, barley, oats, and rye. The rice plant belongs to the genus Oryza. Oryza contains two cultivated species, the Asian Oryza sativa (Figure 6.44) and the West African Oryza glaberrima, each of which includes a multitude of variants. Archaeologists know less about O. sativa than they do about early domesticated wheat, barley, or corn; O. glaberrima is even more of a mystery. East Asia seems to have been the focus for early experimentation and cultivation of O. sativa. On the basis of recent archaeological research in China, archaeologists believe that rice cultivation may have occurred as early as the eighth or ninth millennium B.C. along the middle Yangtze River. By 6500 B.C., rice had become an important component of the diet at sites both to the north and along the lower Yangtze River to the east. In South China, large quantities of rice stalks, grains, and husks were preserved at the waterlogged prehistoric village of Ho-mutu, excavated between 1973 and 1978. The settlement contains houses raised on wooden piles above a lakeshore. Excavations revealed rice in stratigraphic levels extending back to the late sixth and early fifth millennia B.C.
In one place, the rice remains were 25–50 cm (10–20 in) thick, suggesting that a threshing floor may have been preserved. Both wild and domesticated varieties have been identified at Ho-mu-tu. Remains of bottle gourds, water chestnuts, and sour jujubes were recovered, as were acorns and other nuts. The faunal assemblage includes wild deer, elephant, rhinoceros, tiger, and turtle, as well as domesticated water buffalo, dog, and pig. Cord-impressed pottery was found at Ho-mu-tu, and similar ceramic wares have been recovered, along with rice, at other recently excavated sites in South China. Some of these latter sites are earlier, dating to the mid-seventh millennium B.C. Rice, with both wild and domesticated characteristics, also has been recovered at excavated sites in Thailand, where it may pertain to as early as the late fourth millennium B.C. In Southeast Asia, where rising sea levels have covered many Neolithic sites, even earlier evidence of rice domestication may lie underwater. While we still have fewer answers than questions about rice domestication, recent research suggests that rice was domesticated from its wild ancestor in two different areas of eastern Asia. The earliest domestication appears to have occurred in the middle and lower Yangtze River valley in southern China. Rice also may have been independently domesticated farther to the west, on the Indian Peninsula. Nevertheless, the processes surrounding the domestications and the precise timetable for the spread of cultivated rice across southeastern Asia remain largely undeciphered.
Figure 6.44 Oryza sativa, the cultivated Asian species of the rice plant.
jujube A small, edible fruit from an Asian tree of the buckthorn family. The fruit has one seed in the center, somewhat like a cherry.
_ Concept Rice
239
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
Page 240
Site
:
Khok Phanom Di The documentation of the spread of rice into Southeast Asia
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on Khok Phanom Di, see the Internet exercises on your online learning center.
his proposal did not conform to the traditional views of mainland Southeast Asia as a kind of cultural recipient or backwater area that obtained most technological advances from India or China. For Thailand, traditional archaeological reconstructions suggested that the earliest domesticates came from North China after 3000 B.C. During the 1950s, the most relevant archaeological materials for tropical Southeast Asia belonged to the Hoabinhian complex, known from a series of limestone caves and shell middens. Because this complex was not well dated, the presence of cord-marked pottery with edge-ground stone axes in the upper deposits of several Hoabinhian sites was presumed to signal the introduction of agriculture from North China. In the mid-1960s, Chester Gorman began a research project in the uplands
In 1952, geographer Carl Sauer proposed that the cradle for the world’s earliest agriculture should lie in mainland Southeast Asia, rather than in Southwest Asia, as most scholars still believed. Sauer thought that the highly diverse, tropical, and riverine environments of Southeast Asia would have provided a superb locale for the fishing and farming way of life that he surmised would constitute the earliest stage in the transition to agriculture. Like the Russian botanist Nikolai Ivanovich Vavilov before him, Sauer was aware that the wild ancestors for a wide array of modern cultigens (root, tree, and seed crops) had been traced to Southeast Asia. For over a decade, few prehistorians gave serious consideration to Sauer’s ideas because of an absence of information to evaluate them. Moreover,
Figure 6.45 Locations of the Southeast Asian sites mentioned in the text.
VIETNAM
MYANMAR (BURMA)
Hanoi
LAOS Spirit Cave Ban Chiang Non Nok Tha Bay of Bengal
THAILAND Bangkok Khok Phanom Di
South
CAMBODIA
China
Phnom-Penh Saigon
Sea
Gulf of Sites temper A nonplastic material (such as
_ _
sand, shell, or fiber) that is added to clay to improve its workability and to reduce breakage during drying and firing.
240
Thailand
Modern cities
Chapter 6 The Origins of Agriculture
M A L AY S I A
SU
BRUNEI
M
AT R
M A L AY S I A
A
pri31057_ch06.pp198-265.qxd
10/1/09
8:02 AM
of northern Thailand to elucidate the local Hoabinhian hunting-gathering pattern and to document the possible shift to plant cultivation. He began at Spirit Cave, in the rugged, hilly terrain near the Burmese border (Figure 6.45), where excavations revealed a stratigraphic sequence of two cultural levels. The lower deposits, dating from 9000 to 5500 B.C., contained standard Hoabinhian chipped stone artifacts and grinding stones. Polished adzes, cordmarked and burnished pottery, and polished slate knives, similar to tools used to harvest rice in Java today, were found in the upper soil layers, which have been dated by accelerator mass spectrometry (AMS) to the second millennium B.C. The most unexpected findings, however, resulted from the careful study of plant remains. The first archaeologist to carefully sieve the excavated soil from a Hoabinhian site, Gorman found a great variety of seeds, shells, husks, and other plant parts. Food plants such as butternut, almonds, cucumbers, water chestnuts, a few beans, and peppers were recovered, as were remains of the stimulant betel nut, candlenuts (probably used for lighting), and the bottle gourd, a probable container. None of the recovered plant species differs significantly from its wild prototype, indicating that the plants were not domesticated. Nonetheless, a wide variety of plants was intensively utilized, possibly involving the tending or fostering of some species. Significantly, the remains of rice were absent at Spirit Cave. Yet rice has been recovered at other sites broadly contemporaneous with the later occupation of Spirit Cave. At Non Nok Tha, a habitation mound in the lower northeastern part of Thailand, clear impressions and carbonized remnants of rice chaff were found in potsherds. Apparently, the ancient potters, like their modern counterparts, mixed the plant materials into their clay as a fiber temper to improve the workability of the clay and to reduce breakage. In spite of disagreements over whether the rice was domesticated, analyses by Charles Higham, of the University of Otago, New Zealand, of the animal re-
Page 241
mains from Non Nok Tha, and contemporary levels at the more northerly site of Ban Chiang, have revealed an increasing reliance on domesticated animals—cattle, pigs, chickens, and dogs—at these sites. Yet hunting and gathering also continued to be important for subsistence. Cultivated rice has been recovered at Khok Phanom Di (COKE fa-nome DEE), which translates from Thai as “Good Mound.” This coastal site, a 5-ha (12.3-acre) mound that rises 12 m (39 ft) above the surrounding floodplain in a rich estuarine setting near the Gulf of Siam, was occupied between 2000 and 1500 B.C. (Figure 6.46). The first settlers may have been attracted by the broad estuary, its inexhaustible supply of food, abundant potting clay, and a river that facilitated exchange with other regions. Most of the artifacts recovered from the earliest layers relate to fishing and the manufacture of ceramic vessels (Figure 6.47). Recovered bones from animals known to inhabit the nearby mangrove swamps, such as macaques and pigs, reveal hunting activities. The saltwater estuary was not conducive to rice cultivation, so Higham believes that the earliest rice at the site was obtained from inland farming communities by exchange. The coastal inhabitants of Khok Phanom Di also exchanged for exotic ornaments made from shell, ivory, and slate. Hoes and harvesting knives made from shells found in later levels indicate that rice was cultivated late in the occupation of the site. Dog, the only domesticated animal present at the site, also was found in the later levels.
Figure 6.46 The site of Khok Phanom Di is the large, tree-covered mound on the horizon. The site, surrounded by the flat floodplain of the Bang Pakong River, lies 22 km (14 mi) from the present shore of the Gulf of Siam.
Figure 6.47 Excavations at Khok Phanom Di have revealed a 7-m-(23ft)-deep sequence of occupations spanning 500 years from approximately 2000 B.C. Site Khok Phanom Di
241
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Figure 6.48 “The Princess” of Khok Phanom Di. This woman in her midthirties was interred wearing a garment that had been embroidered with more than 120,000 small shell disc beads. She wore a necklace of 1000 I-shaped shell beads and a shell bracelet on her left arm. Two large shell discs were placed on her shoulders.
bier A stand on which a coffin or a
corpse is placed.
_ _
metallurgy The art of separating
metals from their ores.
242
Page 242
The inhabitants of Khok Phanom Di invested significant energy in mortuary ritual. Excavations into the site’s mound exposed over 150 graves. Many of the bodies were covered with red ochre, wrapped in shrouds, or placed on wooden biers. Grave offerings included well-made pottery that was brilliantly burnished and incised with complex designs, shell beads, and bangles fashioned from fish vertebrae. An especially rich grave, sometimes referred to as “The Princess,” was accompanied by thousands of small shell beads, richly ornamented black pottery vessels, a clay anvil, and a potter’s kit that included two burnishing stones (Figure 6.48). The remains lay under a pile of clay cylinders that once had been destined to become pottery vessels. The nature and richness of the burial goods have been interpreted as evidence that the woman was a potter of high status. Yet individuals thought to be her descendants were interred with few grave offerings, suggesting that status may have been achieved through personal endeavor and was not entirely determined at birth or ascribed. The burials at Khok Phanom Di reveal significant social differences in a sedentary community of huntergatherer-cultivators in a rich estuarine setting who exchanged with newly established inland farming communities. Although rice was not originally domesticated at Khok Phanom Di (more likely to the north in South China), the presence of cultivated rice there documents its arrival in Southeast Asia by the second millennium B.C. In recent decades, knowledge about the early Holocene in Southeast Asia has expanded significantly. We have learned that the broad-spectrum Hoabinhian complex extended more deeply into the past than had been previously believed and that plant tending may have contributed importantly to the diet. A more dramatic transition occurred during the past several millennia B.C. with the introduction of cultivated rice, as seen at Khok Phanom Di. At other settled villages, such as Non Nok Tha and Ban Chiang, the inhabitants lived in houses built on
Chapter 6 The Origins of Agriculture
piles or wooden stilts, buried their dead in low mounds, and made socketed spearpoints and adzes as well as adornments of bronze. The appearance of these objects in northeastern Thailand around 1500 B.C. is significant in that they represent some of the world’s earliest known bronzeworking, or metallurgy. More than 800 metal ornaments and weapons were uncovered at Ban Chiang, and many more were found at Non Nok Tha. At these later settlements, rice agriculture and domesticated animals constituted the bulk of the diet. Although much has been learned, many questions remain. We cannot evaluate the importance of indigenous tropical domesticates, such as yam and taro, because these root crops preserve so poorly in the archaeological record. The specifics of rice adoption remain a bit sketchy, although recent research places the original hearth for this key grain in South China. Consequently, although Sauer’s hypothesis for indigenous Southeast Asian plant domestication may eventually hold for certain tropical roots, fruits, and tubers, concerted archaeological studies over the past decades now indicate that the region’s staple grain—rice— was probably introduced into the region from South China.
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 243
Site
:
Guilá Naquitz Cave A preceramic seasonal campsite in Mexico less than a week after beginning a reconnaissance of caves in the area, he found a rockshelter with lots of chipped stone debris and half of a projectile point on the ground surface. This small overhang, named Guilá Naquitz (ghee-LA nah-KEETS), which means “white cliff” in the native Zapotec language of Oaxaca, was subsequently excavated (Figure 6.51). At Guilá Naquitz, Flannery and his field crew carefully peeled away the layers of preceramic occupations dating to 8750–6670 B.C. On the basis of careful retrieval and analysis of the floral, faunal, and artifactual remains found in the cave strata, Flannery suggested that the shelter of Guilá Naquitz was occupied seasonally between August and December by small groups, or microbands, perhaps composed of a series of nuclear families. As in the rest of the Mexican highlands at that time, the ancient inhabitants of Oaxaca were mobile, living in several different
Kent V. Flannery spent the day before Christmas of 1964 in the foothills of the eastern Valley of Oaxaca, in southern Mexico, searching for the best site to study agricultural origins in Mexico. Having just completed several seasons of fieldwork as the faunal analyst on Richard MacNeish’s interdisciplinary team in arid Tehuacán 150 km (90 mi) to the north (see “Tehuacán,” p. 250), Flannery wanted to see what form early agriculture might have taken in a more humid valley with greater farming potential (Figure 6.49). On that December day, Flannery found a preceramic occupation at Cueva Blanca, a site that, he says, more than any other launched his Valley of Oaxaca Human Ecology project (Figure 6.50). On Christmas day, heartened by the previous day’s finds, Flannery began to prepare a research proposal. A year later, with support from the Smithsonian Research Foundation, he returned to Oaxaca. On January 26, 1966,
Texas
0
500 km
SIE RR
SIE
A M
RR
Gulf of Mexico
A
D
A
Figure 6.49 Regions of Mesoamerica mentioned in the text where evidence of early agriculture has been recovered.
RE
M
MEXICO
We have indicated that the research problem we chose for ourselves at Guilá Naquitz was to develop a model that would not only deal with some of the underlying and more universal aspects of early domestication but also tie that process into the specific cultural pattern for the Valley of Oaxaca. —Kent V. Flannery (1986)
D
OR
A
R
IEN
TAMAULIPAS
E
C
TAL
O
C ID EN
TA L
TEHUACÁN VALLEY VALLEY OF OAXACA SIER
R RA M A D R E DEL S U
SIE R
R A DE O AXA
BELIZE CA
CHIAPAS SIE
GUATEMALA RR AM AD RE
PACIFIC OCEAN EL SALVADOR
microband A small family group of
hunter-gatherers.
Site Guilá Naquitz Cave
_
243
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 244
TEXAS
Figure 6.50 The locations of Guilá Naquitz Cave and Cueva Blanca in the Valley of Oaxaca.
Gulf of Mexico
MEXICO
BELIZE
Pacific Ocean Valley of Oaxaca
Oaxaca City
GUATEMALA EL SALVADOR
Guilá Naquitz Cueva Blanca Río Sa
lado
Río
Atoy a
c
Monte Albán
0
5
10 km
N
Piedmont Steep mountains
maguey Any of several species of
arid-environment plants with fleshy leaves that conserve moisture.
_ _
mesquite A tree or shrub of the southwestern United States and Mexico whose beanlike pods are rich in sugar.
244
camps during the course of their yearly activities. The archaeological contents of the cave strata indicate that the inhabitants of Guilá Naquitz consumed a diversity of plant foods, such as acorns and the roasted heart of maguey plants (also called agave, the source of tequila and mescal today), which could have been collected in the thorn forest surrounding the site (Figure 6.52). Other plant foods, such as the seeds of the pod-bearing mesquite tree and hackberries, were brought back to the cave from the flatter grassland below. A small part of the diet, one that increased slightly over time, came from squash and bean plants, which may have been tended or cultivated in the disturbed terrain around the site. The use of wild squash may have been a first step toward eventual domestication. Bruce Smith, of the Smithsonian Institution, has recently restudied squash seeds and fragments (Cucurbita
Chapter 6 The Origins of Agriculture
pepo) recovered from early levels at the site. Accelerator mass spectrometry (AMS) dates taken on some of these remains indicate the presence of domesticated squash just after 8000 B.C. Neither maize cobs nor maize kernels were identified in ancient levels. The variety of nuts, seeds, fruits, and cactus eaten during late summer and early autumn was supplemented by a small amount of venison and rabbit meat. Although deer and rabbit bones were not abundant at Guilá Naquitz, this meat still provided much of the protein consumed in the cave. The subsistence activities of the inhabitants of Guilá Naquitz were rather conservative, changing little over the millennia of intermittent occupations. Perhaps this continuity should not be surprising. Two nutritionists, J. R. K. Robson and J. N. Elias, concluded that the diet of the cave occupants compared favorably with the contemporary diet in the United States, providing similar
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
levels of nutrients and exceeding caloric requirements with a lower intake of food. Whereas seasonally abundant plant foods could have been collected near the cave, stone for making tools was taken from quarries up to 50 km (30 mi) away. This better raw material was used primarily for projectile points, while more local, but less suitable, rock sources were used for more disposable tools, such as scrapers and worked flakes. Besides recording the collection and processing of plant foods and the butchering and consumption of animals, the archaeologists at Guilá Naquitz documented the manufacture of tools, the digging of pits to store acorns, the use of fire pits to prepare food, the processing of plant fibers into strands that were used to make nets (Figure 6.53), and the collection of leaves for bedding in the cave. Years of multidisciplinary study by Flannery and his
Page 245
colleagues illustrate how archaeologists can reconstruct the events of the past into a credible picture of ancient life just before the advent of agriculture in the highlands of ancient Mexico.
Figure 6.52 Large maguey plant in the eastern Valley of Oaxaca. In addition to being a source of food and drink, the plant provides fibers that were used to make ropes, nets, and other woven goods.
Figure 6.51 Excavations in progress at Guilá Naquitz Cave. Many of the species of cacti and thorny bushes outside the cave have seeds or fruits that were eaten by Archaic foragers.
Figure 6.53 A fragment of knotted net from Guilá Naquitz Cave. Such nets were used as bags for transporting food to the cave. Site Guilá Naquitz Cave
245
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 246
Concept
<
Zea mays The mysterious ancestry of corn
Figure 6.54 The great diversity of modern maize, one indication of the tremendous adaptability and genetic plasticity of the plant. These varieties come from many regions of Mexico, and they prosper under different environmental and topographic conditions.
Maize—or corn, as it is familiarly known in the United States—was unknown in Europe before the arrival of Columbus in the Americas in 1492. By that time, the plant was cultivated by native inhabitants over much of the tropical and temperate portions of the Western Hemisphere. The great adaptability and plasticity of maize is evidenced by its position as the second or third most important food plant on earth and its current worldwide distribution (Figure 6.54). Botanical studies, however, indicate that the ancestor of modern corn (Zea mays) was native to southern or western Mexico. Botanists and archaeologists have puzzled over the ancestry of maize.
_ _
246
Chapter 6 The Origins of Agriculture
Domesticated varieties of such cereals as wheat, barley, and rice are structurally nearly identical to living wild species. The principal difference is that, whereas in the domesticated varieties the edible seeds tend to remain fastened on the plant, the wild varieties have shattering inflorescences. As mentioned earlier, shattering is a mechanism by which the seeds of the plant are dispersed naturally. The most recognizable feature of domesticated maize, the massive husked ear, is not present even in wild grasses most closely related to maize. In fact, the nonshattering ear of domestic corn, with its surrounding husks, inhibits seed dispersal, so that without farmers
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 247
Primary tassel
Secondary tassels
Primary tassel
Teosinte spikelet (left) and seeds (right)
Maize cob (left) and kernels (right)
Teosinte plant
to remove and plant kernels from the ear, modern maize could not reproduce for even one or two years. Because of its structure, George Beadle, a renowned plant geneticist, referred to domesticated maize as a “biological monstrosity.” Because of this pronounced difference between the ears of domesticated maize and related wild grasses, the debates over the origins of maize have been more contentious and, until recently, have achieved less consensus than discussions about the origins of most other domesticated seed plants. Although a few experts still hold the view that there was once a wild species of maize (with ears like maize) that has since become extinct (no such plant has ever been found), most participants in the debate have returned to a previously popular position that the ancestor of maize was a variety of teosinte, a giant wild grass so closely related to Zea mays that most botanists place it with corn in the same species (Figure 6.55). Teosinte still grows in the foothills and highlands of Mexico and Guatemala and, in fact, is the only large-seeded, wild, annual grass in the tropical Americas. Unlike maize, teosinte lacks a cob; instead, its seeds are contained in fruitcases. At maturity, teosinte seeds are dispersed through shattering. In other respects, the annual teosinte varieties are very similar
Maize plant
to maize and produce fully fertile offspring when interbred with the domesticated plant. Teosinte and maize have similar tassels, and similar DNA, amino acid, and nutritional compositions. Even examination with a scanning electron microscope cannot distinguish the pollen of annual types of teosinte from small-seeded varieties of corn (such as early domesticated maize). In recent years, new botanical evidence has further clarified the ancestral role of the plant that the Aztecs called teocentli (God’s ear of corn). Biologist John Doebley, of the University of Wisconsin, and botanist Hugh Iltis, emeritus professor at the University of Wisconsin, compared six diverse kinds of teosinte. They found that two annual teosintes were most similar in morphology and other characteristics to maize. Later studies by Doebley determined that the specific proteins from one of these annual teosinte subspecies (Zea mays parviglumis) were indistinguishable from those of maize. The parviglumis teosinte was more similar to maize than any of the other teosintes. Doebley even demonstrated that a geographic cluster of the parviglumis subspecies growing at an elevation of 400–1200 m (1300–3900 ft) along the slopes of the central Balsas River drainage, 250 km (155 mi) west of the Tehuacán Valley, was more biochemically
Figure 6.55 The teosinte and maize plants. Although the kernels are different, pollen from teosinte, the likely ancestor of maize, is almost indistinguishable from the pollen of early corn.
inflorescence The flowering part of
a plant. teosinte (Aztec, teocentli ) A tall annual grass, native to Mexico and Central America, that is the closest relative of maize.
Concept
Zea mays
247
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 248
similar to maize than the other geographic subpopulations of this plant. In fact, recent genetic analysis by Doebley and his colleagues suggests that a small number of single-gene changes could account for the transformation of these annual teosintes into maize. The Balsas drainage, largely unknown archaeologically, may be a promising area for pursuing the initial domestication of maize. To date, archaeological findings have supplemented but not resolved these debates. Archaeological deposits at Guilá Naquitz, dating to the seventh or eighth millennium B.C., contained bean and squash seeds and grains of Zea pollen. Yet we do not know whether the pollen came from maize or teosinte, or how the pollen was transported into the cave. Today, teosinte often grows in the
_ _
same fields with beans, squash, and maize. Teosinte is a weedy pioneer that thrives in disturbed areas such as seasonally wet streambeds and abandoned campsites. Although teosinte can be neither popped nor ground into flour as easily as maize, the wild plant is occasionally eaten as a low-choice or “starvation” food in times of need. Young teosinte ears could have been eaten for their sugary taste. Seasonally, stems of the plant may have been (and are still) chewed, because the pith stores so much sugar. According to Richard MacNeish’s Tehuacán research, the earliest domesticated corn remains appear in cave deposits dating to the end of the sixth millennium B.C. The original dating of these Tehuacán cave strata was accom-
Figure 6.56 Archaeological corncobs recovered during excavations of Tehuacán cave strata document the increasing size of the maize ears through time.
248
Chapter 6 The Origins of Agriculture
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
plished through the conventional radiocarbon analysis of charcoal samples. Two decades ago, the early dates for Tehuacán maize were questioned, based on subsequent direct radiocarbon accelerator dating of the early maize cobs themselves. Accelerator mass spectrometry enables researchers to count individual carbon-14 atoms, rather than relying on the conventional counting of radioactive disintegrations (see “Radiocarbon Dating,” Chapter 4, p. 145). As a result, they can handle much smaller samples. Recent analysis dates the early Tehuacán maize to the mid-third millennium B.C., several thousand years more recent than previously thought. The new AMS dates therefore provide a new date for the arrival of maize in the Tehuacán Valley. Nevertheless, at the same time, pieces of maize cob from strata at Guilá Naquitz have been dated to approximately 4200 B.C., indicating that maize domestication may well have occurred to the west of Tehuacán, possibly in the Balsas.
Page 249
On the basis of these new AMS dates from the valleys of Oaxaca and Tehuacán and other innovative analysis, it is clear that the highland Mesoamerican people who adopted maize were organized in small, seasonally mobile societies. These groups added corn to their way of life, perhaps as early as 10,000 years ago, without radically changing their social or economic behavior. Not surprisingly, this relatively primitive maize does not appear to have become an immediate preceramic dietary staple. Although we cannot directly determine the role of human selection in the evolution of these early ears, later deposits and contexts from the Tehuacán Valley reveal a record of increasing cob size, from the earliest tiny cobs to the much larger and more productive ears (with bigger kernels and more seed rows) grown today (Figures 6.56 and 6.57). The important role of human selection in this latter process is evident.
Figure 6.57 The role of human selection in the domestication of corn is evident in the increasing size of the cob: (a) teosinte, (b) early maize, and (c) modern maize, all drawn to the same scale.
(b)
(a)
(c)
0
5 cm
_ Concept
Zea mays
249
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 250
Site
:
Tehuacán The evolution of early maize
Figure 6.58 Excavated units in Coxcatlán Cave in Tehuacán show its stratigraphy.
Mesoamerica The region consisting of
_ _
central and southern Mexico, Guatemala, Belize, El Salvador, and the western parts of Honduras and Nicaragua that was the focus of complex, hierarchical states at the time of Spanish contact.
250
In the early 1960s, when Richard MacNeish began his fieldwork in the highland Tehuacán (tay-wa-CON) Valley in Puebla, Mexico, little was known about either the preceramic occupation of Mesoamerica or the beginnings of agriculture in the Americas. Before MacNeish’s work, Mesoamerica was theorized to be a hearth of early agriculture. Yet no archaeological evidence existed to support that hypothesis. MacNeish chose to search for the origins of maize (Zea mays) in the relatively small Tehuacán Valley for two reasons. First, because of the region’s
Chapter 6 The Origins of Agriculture
dryness, preservation was unusually good. Preliminary excavations unearthed fragments of basketry and plant materials in limestone cave deposits. Second, MacNeish already had recovered tiny 5000-year-old corncobs in caves in both the northeastern Mexican state of Tamaulipas and the southern state of Chiapas. He reasoned that the earliest domesticated Zea should be still older and would be found in a highland region such as Tehuacán, located between Tamaulipas and Chiapas. MacNeish designed his Tehuacán research to examine two critical ques-
pri31057_ch06.pp198-265.qxd
Date
10/1/09
8:03 AM
Bifacial
Page 251
Unifacial
Ground stone
1000
2000 0
3000
25 cm
Years B.C.
4000
0
25 cm
5000
6000
0
25 cm
7000
8000
9000
10,000
tions: (1) What led to the domestication of maize? (2) How did these changes lay the foundation for later Mesoamerican civilization? He undertook a large survey that located more than 450 prehispanic sites over the 1500 sq km (575 sq mi) of the valley and excavated at a series of 12 cave and open-air deposits, including Coxcatlán Cave (Figure 6.58). Controlled stratigraphic excavations, combined with a large number of radiocarbon dates, enabled MacNeish to reconstruct an unbroken 12,000-year sequence of occupation, at that time the longest recorded in the Americas. For the first time, a picture of early presedentary, preceramic society in Mesoamerica could be sketched, using both artifacts and the plant and animal remains preserved in dry caves of Tehuacán (Figure 6.59). During the preceramic era, according to MacNeish, the few people of the Tehuacán Valley lived in microbands that dispersed periodically. Some camps accommodated only a
single nuclear family, while others sheltered much larger groups. The plant and animal remains recovered from preceramic sites in Tehuacán’s diverse topographic zones led to the recognition that the seasonality of resource availability and the scheduling of resource extraction were critically important in determining the annual regime. More specifically, the early inhabitants of Tehuacán scheduled their seasonal movements across the highland region, from the riverbanks to the foothills to the mountains, to coincide with the periodic availability of local plant and animal species. For most of the preceramic era, such game as rabbits and deer supplemented plants in the diet. During the May–October rainy season, edible plants were more abundant, and a diversity of seeds, cactus fruits, and berries were exploited, in addition to the bountiful seedpods of the mesquite tree. Rabbits, rodents, lizards, and other small animals were consumed at this time of year, when the size of human groups
Figure 6.59 Changes in the flaked and ground stone industry in Tehuacán, 10,000–1000 B.C.
From a personal point of view, and closely connected with satisfying my innate curiosity, has been the thrill of discovery of art and artifact and the fact that these objects have remained unknown to anyone until my little trowel or paintbrush uncovered them—often after long periods of searching for these new and thrilling finds. I have been doing archaeology now for over forty years but the thrill is still there and I feel that this thrill should happen to anyone and everyone. —Richard MacNeish (1978)
seasonality The changing availability of resources according to the different seasons of the year. scheduling The process of arranging the extraction of resources according to their availability and the demands of competing subsistence activities.
Site Tehuacán
251
_
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 252
Figure 6.60 Changes in diet and farming strategies in the Tehuacán Valley, 6000 B.C.–A.D. 1000. Wild plants
Cultivated squash and beans
Cultivated corn
Cultivated fruit and seed plants Domesticated dog and turkey
Big and small game
6000 BC
5000 BC
4000 BC
3000 BC
2000 BC
1000 BC
0
AD
1000
AD
1000
Wild plants Agriculture
Hydro-horticulture
Irrigation agriculture
Orchards
Meat
6000 BC
_ _
252
Chapter 6 The Origins of Agriculture
5000 BC
4000 BC
3000 BC
2000 BC
1000 BC
0
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
was generally larger. Although some fruits were still available in the early part of the dry season (November and December), cactus leaves and deer apparently were the staples during the dry spells that lasted from January to April. Although this way of life persisted for at least 6400 years, from about 8000 to 1600 B.C., several important dietary changes did take place. A wild ancestor of the domesticated squash was used by roughly 8000 years ago, probably as a container or for its protein-rich seeds. Thereafter during this 6400-year period, domesticated varieties of squash and maize appeared. These early maize ears were small (about the size of an index finger) and contained no more than eight rows of kernels (see Figure 6.56). These early domesticated plants did not immediately provide a large portion of the diet, which still was based primarily on wild plants and animals. Thus these initial experiments in plant domestication occurred among a population that was largely mobile and remained so for thousands of years. Somewhat enigmatically, the bone chemistry assay of human bones from the Tehuacán excavations provides a rather different picture for the period after 5000 B.C. (see “Bone Chemistry and Prehistoric Subsistence,” Chapter 5, p. 176). These studies indicate a smaller amount of meat in the diet than is seen in the archaeological deposits, as well as a greater role for either early cultigens or wild setaria grass, the seeds of which were present but not recovered in abundance in the archaeological record (Figure 6.60). Yet the relative importance of meat inferred from archaeological deposits is not very surprising, given that bone generally preserves better in ancient deposits than do smaller plant materials. In addition to the gradual increase in the overall proportion of both wild and domesticated plant foods in the diet, the Tehuacán sequence reveals an increase in population and a decrease in residential mobility. Based on the
Page 253
size and number of sites known from the preceramic phases, the total population density for the Tehuacán Valley, though low, appears to have increased severalfold during this period. Although the earliest sedentary villages in Tehuacán did not occur until 4000–3000 years ago, the length of site occupation increased, and the size of sites grew during the preceding millennia. A single circular pithouse, the earliest known in Mesoamerica, was found in a 5000year-old level at an open-air site in the region. Later preceramic occupations also tend to have more storage features. There is little question that MacNeish’s Tehuacán research has revolutionized our knowledge of early Mesoamerica, as well as our understanding of the diversity of situations in which early agriculture developed. In the Tehuacán Valley, the first experiments toward plant domestication occurred among people who remained residentially mobile for thousands of years, a sequence that is very different from what has been long known for early farming in Southwest Asia. The wild ancestors of the major Mesoamerican cultigens—maize, beans, and squash—were all highland plants. Thus it is not surprising that the earliest archaeological evidence for Mesoamerican agriculture has been found in highland valleys such as Tehuacán and Oaxaca. The dry caves in these upland valleys are recognized for their superb archaeological preservation. Yet some of Mesoamerica’s earliest sedentary villages were established in the lowlands, where the highland cultigens eventually were incorporated into a coastal subsistence economy that also included marine resources and lowland plants.
How and why has AMS dating of archaeological samples revolutionized our perspective on early agriculture?
y
setaria A wild grass with edible seeds.
_ Site Tehuacán
253
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 254
Site
:
Guitarrero Cave The origins of domestication in the high Andes A century before the arrival of the Spanish in the sixteenth century A.D., prehispanic America’s largest empire was in place along the western side of South America (see Chapter 9). This political domain stretched from the Inca capital of Cuzco, in Peru’s southern highlands, down into Chile and northwestern Argentina and up through Ecuador. It encompassed high Andean mountain
Figure 6.61 The location of Guitarrero Cave in the mountains of northern Peru.
GUIA
NA
HIG
ATLANTIC OCEAN HL
AN
DS
Amazon River
A N
Amazon
S
Basin
OCEAN
I AN ZIL
BRA
0
_ _
254
DS
A N D E S
PACIFIC
Atacama Desert
Lake Titicaca
N LA
HI
GH
E
Desert
D
Sechura
Guitarrero Cave
Chapter 6 The Origins of Agriculture
1000 km
slopes, Pacific coastal deserts, and the western fringes of the Amazonian tropics. The Inca established a network of roads and trails to move people and goods across these diverse topographic zones. Thousands of years before the beginnings of Inca expansion, the earliest steps toward agriculture were taken in this part of South America, where the high Andes are sandwiched by a coastal desert and a tropical rain forest. Although roads were not yet built, communication between these environments was critical to the beginnings of Andean domestication. By 10,000–8000 years ago, Amazonian plants had been introduced into the Andes. After 6000 B.C., morphologically wild plants and animals from the rain forest and mountains were present at sites established along the Pacific. Yet the earliest indications for cultivation and domestication appear in the mountains. As mentioned in the earlier discussion of Monte Verde, human groups first arrived in South America as mobile hunters and gatherers, using stone implements and eating a variety of foods (see “Monte Verde,” Chapter 4, p. 154). They spread rapidly across the continent. Although earlier sites have been reported, archaeologists have a much more complete understanding of the past 10,000–12,000 years in the Andean highlands. One of the important sites for the early period is Guitarrero (gheeta-RARE-o) Cave, a large natural rockshelter at 2580 m (8500 ft) above sea level in the mountains of northern Peru (Figure 6.61). First occupied more than 10,000 years ago, Guitarrero Cave served as a campsite for thousands of years, accumulating a valuable record of the beginnings of domestication in the Andes (Figure 6.62). Thomas F. Lynch, of Cornell University, directed excavations in the
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 255
Figure 6.62 A view of the Andes from Guitarrero Cave, a large natural rockshelter situated 2.5 km above sea level in the mountains of northern Peru. The site was first used by Native Americans more than 10,000 years ago and contains a valuable record of the beginnings of domestication in the South American highlands.
cave during the late 1960s (Figure 6.63). The dry highland cave environment preserved many organic materials, and polished bone knives, fragments of gourd bowls, cordage, basketry, and textiles were recovered (Figure 6.64). C. Earle Smith noted that the total bulk of the inedible fibrous plants was roughly equivalent to that of the food plant remains at the site, suggesting that the Andean utilization of fibers and textiles had a very early origin (Figure 6.65). The basic twining, or
finger-weaving, technique, used to create many of the textiles at Guitarrero Cave, is clearly an important step toward the elaborate techniques used for later prehispanic Andean fabrics. Organic remains also revealed continuity in another important Andean cultural pattern: communication and exchange between different environmental areas. Guitarrero Cave, on the western slope of the Andes, sits in the middle of three principal environmental zones. The high-altitude shelter
_ Site Guitarrero Cave
255
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 256
Figure 6.63 Excavating the early Archaic deposits at Guitarrero Cave.
0
_ _
1 cm
Figure 6.64 Early Archaic flint scraper wrapped in deer hide and cord recovered from Guitarrero Cave.
256
Chapter 6 The Origins of Agriculture
faces the exceedingly dry, narrow Pacific coastal zone. To the east, over the Andes, are the wetter, tropical eastern slopes, or montaña zone, which grades down to the Amazon jungle. The presence of lima beans, a plant native to the Amazon Basin, in archaeological levels dated to roughly 8000 years ago, reveals both the antiquity of pan-Andean connections, at least in an indirect, down-the-line fashion (passing from person to person, group to group over long distances), and the origins of cultivation in South America. As with the lima bean, many wild ancestors of the major South American domesticated plants, including manioc, peanut, guava, and coca, were native to the eastern side of the Andes. Yet because of the poor archaeological preservation in the montaña zone, we know little about the first stages in the process of cultivation and domestication. Other root crops, such as the potato, are indigenous to the mountain
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 257
Figure 6.65 Basket fragments recovered from Guitarrero Cave.
0
zone. Unfortunately, efforts to learn the origins of root crops are hampered because these plants reproduce asexually (see “The First Farmers,” p. 199). Ten thousand years ago, Andean tubers and rhizomes were prime sources of carbohydrates for the occupants of Guitarrero Cave. Yet these plants may have been wild, collected from higher elevations. As in highland Mesoamerica, the first steps toward cultivation and domestication in the Andean highlands resulted in few immediate changes in society. In both regions, the preceramic era was characterized for thousands of years by continued mobility and a rather resilient diet. At Guitarrero Cave, the earliest inhabitants relied for generations on a variety of tubers, rhizomes, and squash for carbohydrates; several kinds of beans for plant protein; and wild fruits, along with a variety of chili peppers, for minerals and vitamins. Analysis of the faunal remains from Guitarrero by Elizabeth Wing, of the University of Florida, indicates that deer, camelids, rabbits, and a range of small animal and bird species were hunted. Wing observed a steady decline in the number of deer bones and an increase in those of camelids through time, a pattern also seen at other An-
1 cm
dean sites from this time and particularly at sites in the higher grassland, or puna, region. In the Andes, the increase in camelid remains is only part of the evidence for manipulation and eventual domestication (for both meat and wool) of the larger llama and the smaller alpaca. The larger llama also served as a pack animal, carrying goods across the Andes. Another Andean species, the guinea pig, also was domesticated for meat but was not abundant at Guitarrero Cave.
montaña (Spanish) Mountain, specifically referring to the wet, tropical slopes of the Amazonian Andes. tuber A fleshy, usually oblong or rounded outgrowth (such as the potato) of a subterranean stem or root of a plant. rhizome An edible, rootlike subter-
ranean plant stem. puna (Spanish) High grassland
plateaus in the Peruvian Andes.
_ Site Guitarrero Cave
257
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 258
Concept
<
Agriculture in Native North America Indigenous plant domestication before the spread of maize By the time Europeans arrived in the Americas, the maize plant had spread far to the north and south. Its arrival in Central and South America followed the plant’s initial domestication in highland Mesoamerica. Yet until recently, the exact timing of the plant’s introduction to the south had been a matter of considerable conjecture. The recent AMS dating of maize starch grains that were extracted from charred residue on ceramic cooking vessels has now placed domesticated maize in Ecuador by around 3000 B.C. In South America, maize supplemented a range of other agricultural plants that were indigenously domesticated. The history of maize in the Americas north of Mexico is somewhat clearer. Maize reached the Southwest by
_ _
Figure 6.66 The seeds of domesticated marsh elder, sunflower, and squash (top to bottom, on the right) are larger than those of their wild ancestors (on the left). Larger seeds germinate faster and produce plants that grow more quickly, resulting in greater quantities of food. Selection in the past for larger seeds was an important part of the domestication process of these seed plants.
258
Chapter 6 The Origins of Agriculture
2100 B.C., and from there it apparently was carried across the Great Plains and arrived in the Eastern Woodlands by approximately 200 B.C.–A.D. 200. Although the timing of maize’s North American arrival is now better established, the role of the exotic domesticates in the beginnings of North American agriculture is still a matter of discussion. Formerly, most scholars postulated that agriculture in eastern North America began in response to the diffusion of exotic domesticated plants (specifically, squash) from the south, but that view is beginning to be challenged in the face of mounting archaeological and genetic evidence. Based on excavations at a range of sites in eastern North America, from Ohio to Arkansas, archaeologists have found that native North American seed plants were domesticated in riverine floodplain settings as early as 2500–1500 B.C. The plants include chenopod (Chenopodium berlandieri, also known as goosefoot or chia), sunflower (Helianthus annuus), and marsh elder (Iva annua, var. macrocarpa). In each case, botanical studies have shown that the domesticated varieties have larger (Figure 6.66) and morphologically different seeds than the wild ancestors (Figure 6.67). As in Mesoamerica and South America, these domesticated seed plants were incorporated into the diet of mobile peoples, who also exploited a range of wild foods. Although AMS dates on the maize found in eastern North American caves have consistently placed the arrival of this Mesoamerican plant well after the domestication of native North American seed plants, the picture is a bit fuzzier for squash (Cucurbita pepo). Small carbonized pieces of squash rind were recovered in flotation samples collected at river valley sites in Illinois
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
and Kentucky. These fragments yielded direct AMS dates of roughly 5000–3500 B.C. Though C. pepo does include some wild gourds, it is best known for its rich variety of domesticated squashes and pumpkins. Researchers also documented its domestication in highland Mexico around 8000–7000 B.C. The archaeological contexts that produced these early pieces of rind were outside the current range of the Texas wild gourd, another Cucurbita species that grows in eastern Texas. An obvious initial conclusion was that domesticated squash was introduced to eastern North America, along with the notion of agriculture, well before the cultivation of native plants. This first scenario has been challenged by an alternative viewpoint suggesting that C. pepo was independently domesticated twice from different ancestral populations of wild gourds. It was domesticated in Mexico and then later in eastern North America around 2500 B.C. (around the same time the indigenous seed plants were domesticated). All the early eastern North American Cucurbita rind samples are thin and likely came from wild gourds. Early seeds from these plants also are small, within the range of wild gourds. The earliest morphological marker, larger seed size, of Cucurbita domestication in eastern North America did not appear until 2500–2300 B.C. These interpretations are supported by the findings of botanist Deena DeckerWalters, of Texas A&M University. Her biochemical analyses of C. pepo have revealed that the species is composed of two developmental lineages, one from Mexico and the other from eastern North America. This view has recently been supported by the discovery of wild squash in the Ozarks of Missouri and Arkansas. This Ozark gourd occupies the same floodplain niche inhabited by two of the seed plants (goosefoot and marsh elder) that were indigenously domesticated in the East. On the basis of these recent findings, researchers believe that the early squash rind found in eastern North
Page 259
Figure 6.67 The compaction of seeds is a good indicator of domestication. In the wild Chenopodium plant known as lamb’squarters, on the right, seeds are distributed in numerous small clusters. On the left, in contrast, the seeds of a domesticated Chenopodium plant are tightly compacted at the top of the main stem.
America was probably from a native plant. Consequently, agriculture appears to have occurred indigenously in this North American region, well before the diffusion of the concept or any plant from Mexico.
_ Concept Agriculture in Native North America
259
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 260
Concept
<
Breast-Feeding and Birth Spacing Ideas about post-Pleistocene demographic growth
The question must be raised. Why farm? Why give up the 20-h [hour] work week and the fun of hunting in order to toil in the sun? Why work harder for food less nutritious and a supply more capricious? Why invite famine, plague, pestilence and crowded living conditions? Why abandon the Golden Age and take up the burden? —Jack Harlan (1992)
weaning The process of transferring the young from dependence on its mother’s milk to other forms of nourishment. lactational amenorrhea The sup-
pression of ovulation and menstruation during breast-feeding.
It is indisputable that the beginnings of farming and especially sedentary life were associated temporally with an increased rate of population growth. Because the reasons for that rapid growth are difficult to understand through the archaeological record alone, archaeologists and physical anthropologists have used demographic and physiological analyses of contemporary peoples to gain insight into factors that may have prompted that transition. Like all significant demographic shifts, the post-Pleistocene changes theoretically could have resulted from changes in migration, mortality, or fertility. Although migration may have become an important factor in certain later agricultural regions, such as Neolithic Europe, it alone cannot account for the rapid growth evidenced in much of the world following food production and sedentism. Although postPleistocene mortality rates are difficult to evaluate, several conditions may have promoted increased mortality following agriculture and sedentism. In most instances, agricultural diets were less balanced than hunting-andgathering diets. In several regions, tooth decay is found in adult burials only after the transition to agriculture. Animal domestication brought people and animals closer and may have led to the genesis of contagious human diseases, apparently beginning as animal viruses or infections. Sedentary populations and higher population densities also may have increased human susceptibility to certain contagious diseases. Thus although longer life spans may have contributed to post-Pleistocene demographic shifts, fertility changes seem to be a more likely impetus.
_ _
260
Chapter 6 The Origins of Agriculture
Demographic studies of contemporary human groups usually, though not always, indicate that settled agriculturalists have higher fertility rates than mobile hunter-gatherers. In human populations numerous factors— including age at marriage, the length of a woman’s reproductive period, birth spacing, coital frequency, and the importance of contraception—can affect the number of births each female has during her lifetime. In contemporary studies of recently settled huntergatherers, different sets of factors seem to contribute to the frequently noted rise in fertility with the introduction of agriculture and sedentism. Despite those differences, shortened nursing time and earlier child weaning often do occur among recently settled foraging populations. Shorter periods of breast-feeding prompt hormonal changes that increase the chances of conception, so spacing between births often declines when women adopt a shorter nursing period. Alternatively, nursing may stimulate a woman’s hormones in a manner that suppresses ovulation and menstruation, a phenomenon known as lactational amenorrhea. Prolonged breast-feeding is therefore one factor associated with longer spacing between births. The Gainj of highland Papua New Guinea, a horticultural village population, have a birth-spacing pattern, without contraception, that is more similar to that of contemporary huntergatherers than to that of other farmers. Gainj women breast-feed their infants for an extended period, more than 3 years on average. In part, the Gainj women breast-feed because they lack adequate weaning foods (a characteristic of many hunter-gatherers as well).
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 261
Figure 6.68 Millet was an early domesticated plant in China, and various subspecies were domesticated or used in Africa and Europe.
The Gainj agricultural staples—sweet potatoes, yams, and taro—are not easy to digest, nor are they the compact sources of nutrients required for weaning, since infants consume only small amounts of food daily. In contrast, the Neolithic staples, which included cereal grains in most regions, of Southwest Asia provided considerably better alternatives and supplements to mother’s milk than are available to the agricultural Gainj. Not only were the high carbohydrate, easyto-digest grains (such as wheat, barley, and millet) available (Figure 6.68), but milk from sheep, goats, and cattle also
could be fed to infants. The increased rate of demographic growth in postPleistocene Southwest Asia could have been linked to changing patterns of weaning, nursing, and hence birth spacing.
_ Concept Breast-Feeding and Birth Spacing
261
pri31057_ch06.pp198-265.qxd
< ( : < V : y < p : W
10/1/09
8:03 AM
Images and Ideas The Spread of Agriculture The success and consequences of food production
D
The question “why farm?” strikes most of us modern humans as silly. Of course it is better to grow wheat and cows than to forage for roots and snails. But in reality, that perspective is flawed by hindsight. —Jared Diamond (2002)
_ _
262
Page 262
ifferent concepts and methods are used by researchers who study the Paleolithic and those who investigate agricultural societies. Paleolithic archaeologists tend to excavate in natural stratigraphic levels, often in caves, paying careful attention to the distributions of bone and stone materials. Researchers focused on later agricultural peoples concentrate more on the stratigraphic levels revealed in the architectural remains of past structures, such as temples and houses, and they must analyze pottery and metalwork as well as stone and bone. Because pre-Neolithic sites have fewer artifacts and are often deeply buried, systematic regional surveys, which find and map archaeological remains visible on the land surface, are rarely as practical or as useful for the specialist in the Pleistocene as they are for many archaeologists who study later periods. These differences in archaeological perceptions and practices reflect real changes in the nature of the archaeological record that began 10,000–3000 years ago in many (though not all) areas of the world. In most regions, Paleolithic sites tend to be small, thin scatters of lithic materials, reflecting generally lower populations and the tendency for occupations to involve fewer people over shorter periods of time. The low artifactual densities and the general absence of substantial structures at prefarming sites also suggest greater residential mobility. In many regions, the presence of residential and civic architecture and cemeteries at archaeological sites is unique to prehistoric farming societies. In addition, whereas the artifacts of the Paleolithic—spearpoints, knives, and scrapers—tended to be largely for capturing energy (food), “facilities,” or materials (stone bowls, ceramic containers) and features (pits) to store energy became much more important in later eras. Kent Flannery aptly noted that there are more storage facilities at ‘Ain Mallaha alone than are known from all earlier Southwest Asian sites, indicating a sudden transition to residential stability, with its implied social and economic adjustments. The archaeological record reveals that a more sedentary way of life, plant and animal domestication, and pottery were not adopted simultaneously, nor did those changes occur in a single or uniform sequence in all regions. For example, sedentary villages preceded any evidence for plant domestication in Southwest Asia, whereas the domestication of maize, beans, and squash occurred before the earliest Mesoamerican villages. Furthermore, although the Mesoamerican combination of beans and corn provides a complete protein source, as well as adequate calories, some of the other staple grains (barley and rice) are high in carbohydrates but low in protein. In addition, the domestication of animals clearly was a much more significant part of the Neolithic transition in the Old World than in the Americas. Many of the regions of early, indigenous domestication were relatively arid (Southwest Asia and highland Mexico). Yet that was not the case in either South China or the riverine settings of eastern North America. Given the varied climatic, demographic, and cultural conditions of the early Holocene, the successful adoption and rapid spread of food production clearly was a widespread phenomenon. In regions where indigenous resources were not domesticated, exotic cultigens and animals often were quickly introduced and adopted. Once domesticated, wheat and barley were transmitted rapidly to the Nile Valley, many parts of Europe, and North China as a supplement to millet. In the river valleys of eastern North America, maize from Mexico
Chapter 6 The Origins of Agriculture
pri31057_ch06.pp198-265.qxd
10/1/09
8:03 AM
Page 263
Figure 6.69 Some major African food crops (left to right): yam, finger millet, and sorghum.
was incorporated into local agricultural complexes that included different combinations of oily seed plants, such as marsh elder and goosefoot. The Neolithic transition apparently was an even more complex mosaic in sub-Saharan Africa. Imported sheep and goats were introduced from Southwest Asia into diverse economies that locally domesticated more than a dozen plant species, including finger millet (Ethiopia and northern Uganda), sorghum (Lake Chad to the Nile), African rice (middle delta of the Niger), tiny-seeded teff (Ethiopia), and yams (West Africa) (Figure 6.69). Whereas domesticated sheep and goats clearly were foreign, the origins of domesticated cattle in sub-Saharan Africa are less certain; they either were exotic or were domesticated independently in northern Africa, where they were present by the fifth millennium B.C. As in Japan, where ceramics preceded food production, pottery vessels were found at sites occupied by semisedentary fisher-foragers from the fifth and sixth millennia B.C. in the Sudan, as well as along the margins of now-dry lakebeds south of the Sahara Desert. The beginnings of cultivation also are being unraveled in the islands of the Pacific, where the transition from foraging to farming clearly was more gradual, with few stark or immediate shifts in lifeway. On the basis of recent studies by Tim Denham, of Flinders University in Adelaide, Australia, and his colleagues at the Kuk Swamp site in the highlands of New Guinea, it appears possible that the banana may have been cultivated as early as 7000 years ago. From a global perspective, food production emerged over the past 12,000 years, broadly coincident with major cultural changes that have shaped the course of recent human history, along with significant climatic and sea-level shifts. Just as the Neolithic creates a divide in the archaeological record and among archaeologists, the beginnings of domestication are generally (although not always) linked in time with more permanent or sedentary communities, changing social and political relationships, larger and denser populations, new technologies, and shifting networks of exchange and communication. The complexity of these relationships and their diversity from region to region make it difficult to decipher the exact causes for prehistoric changes. Yet archaeologists have developed some ideas by studying contemporary peoples, particularly those who have recently shifted from a mobile hunting-and-gathering way of life to more sedentary and agricultural pursuits. _ Images and Ideas The Spread of Agriculture
263
4 2 0
10
8
10/1/09
8:03 AM
Irrigation farming
6
Early dry farming
People per sq km
8
Hunting and gathering
pri31057_ch06.pp198-265.qxd
6
4
Thousands of years B.C.
Figure 6.70 Population densities associated with different subsistence strategies.
Soil erosion, desertification, water pollution and soil degradation are intimately related to agriculture in terms of both the past and the present; they are not phenomena of the twentieth century. —Antoinette Mannion (1999)
p
Page 264
During the Pleistocene, human populations grew at a slow pace (Figure 6.70). If modern hunter-gatherer populations are a reliable guide, a few individuals may have separated from parent groups when resources were exhausted or when disputes became common. In larger campsites, the latter could have been a potential problem, because as the population grows, the number of interactions between individuals increases even more rapidly. By the end of the Pleistocene, with the peopling of most of the continents (albeit at low densities), fissioning would have become less of an option for many groups. This pattern may partially account for observed late Pleistocene–early Holocene increases in local population densities, residential stability, greater reliance on lower-quality foods, and decreases in social group territory sizes. Over the past 10,000 years, the earth’s population has doubled 10 times, from less than 10 million people to more than 6.7 billion. Following the end of the Pleistocene, the more rapid rates of demographic growth may have related in part to changing patterns of child rearing and diet. Recently settled huntergatherers often witness a reduction in birth spacing. A mobile way of life limits the number of infants a family group can transport and hence care for at any one time. Intensive exercise and prolonged breast-feeding have been suggested as factors that temporarily diminish a woman’s fertility. With increased sedentism, storage, and domestication, both the frequency of intensive female activity and the length of the nursing cycle may have decreased. Contemporary huntergatherers tend to breast-feed for 2–6 years because of the frequent absence or unreliability of soft weaning foods. The storable, staple cereal grains provide such an alternative food to mother’s milk. In certain regions, animal milk also could be substituted. More productive and storable food resources may have permitted larger communities and denser populations, yet a series of organizational changes often occurred at roughly the same time. For long-term maintenance and survival, larger communities would have required new mechanisms for integration, cooperation, the resolution of disputes, and decision making. Kin relationships are severely tested when decisions must be made for groups of several hundred. Increased evidence for burials, ritual objects (e.g., figurines), nonresidential structures, more formal patterns of exchange, and in some cases more unequal access to goods and labor would seem to signal these very significant changes in social and political relationships. Some of these new organizational forms were hierarchical, with more permanent, formal leadership roles instituted above the remainder of the population. Such leaders or decision makers, in turn, may have fostered greater concentrations of resources and labor, leading to intensified production and even larger communities. The transition to agriculture and sedentism occurred at the onset of a rapid succession of changes that have culminated in our modern world. The pace of these recent changes is truly remarkable when viewed from the perspective of all of human history. Although neither domestication nor sedentism alone is necessary and sufficient to explain the formation of early cities and ancient states, both permanent communities and food production are critical elements of those very significant, later developments.
DISCUSSION QUESTIONS 1. Where are the six cradles of agriculture around the world? 2. What are the most important of the early domesticated plants and animals?
w w w. m h h e . c o m / p r i c e i p 6 e
_ _
For more review material and study questions, see the self-quizzes on your online learning center.
264
3. Discuss the different kinds of natural environments in which early domestication took place. 4. Australia is one of the few places within the limits of cultivation where agriculture did not spread. Why might that be?
Chapter 6 The Origins of Agriculture
pri31057_ch06.pp198-265.qxd
10/15/09
10:53 AM
Page 265
5. Compare the process of change from foraging to farming in Southwest Asia and highland Mesoamerica. 6. Compare and contrast the role of social, environmental, and demographic factors in the emergence of agriculture in different areas of the world. 7. Discuss how and why human lives tended to change with the advent of sedentary settlements. 8. What are the long-term implications of the origins of agriculture in regard to contemporary concerns about the environment, natural resources, and ecological diversity?
z
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e. Bellwood, P. 2005. First farmers: The origins of agricultural societies. Malden, MA: Blackwell. A comparative study of the origins and dispersal of agricultural communities. Cauvin, J. 2000. The birth of the gods and the origins of agriculture. Cambridge: Cambridge University Press. A new perspective on the beginnings of domestication, focusing on conceptual changes, translated from the French. Clutton-Brock, J. 1999. A natural history of domesticated animals. Cambridge: Cambridge University Press. A detailed compendium of recent information on the domestication of animals. Cohen, M. 1977. The food crisis in prehistory. New Haven, CT: Yale University Press. A comprehensive indictment of global population increase as the cause of the domestication of plants and animals. Flannery, K. V., ed. 1986. Guilá Naquitz: Archaic foraging and early agriculture in Oaxaca, Mexico. New York: Academic Press. A detailed report on an important preceramic settlement in highland Oaxaca in the distinctive style of Flannery. Harris, D. R., and G. C. Hillman, eds. 1989. Foraging and farming: The evolution of plant exploitation. London: Unwin Hyman. A volume of papers on early agriculture, presenting evidence from around the world.
Hodder, I. 2006. The leopard’s tale: Revealing the mysteries of Çatalhöyük. New York: Thames & Hudson. A first-hand description of the recent finds at Çatalhöyük. Price, T. D., and A. B. Gebauer, eds. 1995. Last hunters–first farmers: New perspectives on the transition to agriculture. Santa Fe, NM: School for American Research. A series of papers dealing specifically with the question of cause in regard to the origins and spread of agriculture. Smith, B. D. 1998. The emergence of agriculture. New York: Scientific American Library. A well-illustrated and up-todate summary of agricultural beginnings worldwide. Zeder, M. A., D. Bradley, E. Emshwiller, and B. D. Smith, eds. 2006. Documenting domestication: New genetic and archaeological paradigms. Berkeley: University of California Press. A collection of scientific papers dealing with the most recent evidence for early domestication. Zohary, D., and M. Hopf. 2001. Domestication of plants in the Old World: The origin and spread of cultivated plants in West Asia, Europe, and the Nile Valley, 3rd ed. Oxford: Oxford University Press. A botanical tour of the first domesticated plants in Asia, Europe, and Africa.
_ Suggested Readings
265
pri31057_ch07.pp266-319.qxd
C
10/1/09
H
8:06 AM
A
P
Page 266
T
7
E
R
S
E
V
E
N
Figure 7.1 Monks Mound towers over the surrounding terrain at the site of Cahokia in southern Illinois. This immense earthen mound, which was constructed in stages, beginning in the tenth century A.D., is the largest prehistoric structure in what was to later become the United States.
_ _
pri31057_ch07.pp266-319.qxd
10/1/09
8:06 AM
Page 267
Native North Americans Introduction
< ( : < V : y < p : W
w w w. m h h e . c o m / p r i c e i p 6 e
The Diversity of Native American Life Hunter-gatherers, farmers, and chiefs
T
he towering Monks Mound that dominates the ancient site of Cahokia in southern Illinois was by far the largest earthen feature built north of Mexico by Native American peoples (Figure 7.1). Yet it was only one of many such mounds and earthworks that dotted the landscape of eastern North America before the arrival of Europeans roughly 500 years ago. Nevertheless, by the time of the American Revolution, these landscape features had been largely destroyed or were ignored by most Euro-Americans, and those colonists who did take note questioned their historical connection to the indigenous populace who had inhabited North America for millennia. Interestingly, it was Thomas Jefferson, the third president of the United States, who played an early role in reestablishing the evidential link between North American earthworks and indigenous Native American peoples. Jefferson’s late-eighteenth-century exploration of an earthen mound near Charlottesville, Virginia, was the first stratigraphic study of a North American earthen feature, and he accurately described how that feature served as the repository for a significant number of burials. Although Jefferson did not immediately connect the mound to Native Americans, he eventually did so. Through this scientific contribution, Thomas Jefferson can be considered a father of American archaeology as well as the principal father of the Declaration of Independence. On the twelfth of October in 1492, when Columbus landed on a small island in the Caribbean, roughly 40 million Native Americans resided in the Western Hemisphere. These peoples spoke about 400 mutually unintelligible languages, compared with the more than 2500 languages that were spoken in the Eastern Hemisphere. Yet the diversity of the “spoken tongues” is impressive, considering that the landmass of North and South America is less than half the size of Eurasia and Africa. One important factor influencing linguistic divergence is time, and hominids have lived for only thousands of years in the Western Hemisphere compared with millions of years in the Eastern Hemisphere. At the time of European contact, the populations of the Western Hemisphere were diverse in many ways other than language. Some peoples were incorporated into the great empire of the Inca in the Peruvian Andes; others paid tribute to the rulers of Aztec Tenochtitlán in the Basin of Mexico. Yet not all Native American peoples were as socially and
267
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
CHAPTER OUTLINE
Introduction The Diversity of Native American Life 267 SITE Poverty Point SITE Hopewell
272
276
CONCEPT The Archaeology of Exchange 281 SITE Cahokia
283
CONCEPT Monumental Architecture 287 SITE Moundville
289
CONCEPT Grave Offerings SITE The Draper Site SITE Snaketown
293
295
300
CONCEPT Studying Community
Plan at Snaketown 305 SITE Chaco Canyon SITE Ozette
306
311
CONCEPT Chiefs
315
Images and Ideas
The Clash of Worlds 317
_
pri31057_ch07.pp266-319.qxd
10/1/09
8:06 AM
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the Native American Graves Protection and Repatriation Act, see the Internet exercises on your online learning center.
prehistory In general, the human
past; specifically, the time before the appearance of written records. ethnohistory The study of ancient
_ _
(often non-Western) cultures using evidence from documentary sources and oral traditions, and often supplemented with archaeological data.
268
Page 268
economically stratified as the Inca and the Aztecs. Although some Mississippian lords in eastern North America may have received tribute from surrounding populations, the sociopolitical formations in what is now the eastern United States were not as complex as those farther south. More egalitarian hunter-gatherers lived in the Arctic and sub-Arctic, on the Plains, in desert areas of the western United States and Mexico, and at the tip of South America. The Americas were first colonized by hunter-gatherers who either crossed the broad Bering land bridge or traveled from Asia to the Americas by boat. Initially, their way of life was highly mobile. In western North America, the colonists relied in part on the hunting of large herd animals; the subsistence regime in eastern North America was generally more diversified. In much of the Americas, a long Archaic period followed the millennia (the Paleoindian period) during which several species of large herd animals that were important food sources became extinct. (See “Beringia” and “Lindenmeier,” in Chapter 4, for a discussion of the Paleoindian period and “Carrier Hills,” in Chapter 5, for a discussion of the Archaic.) During the Archaic, regional populations diversified their subsistence pursuits, each concentrating on a distinct set of local resources, hunting a wide range of animals, and gathering a variety of wild plants. In the later Archaic period, cultigens were added to the diet. In midwestern North America, local seed plants were domesticated during the Archaic before the arrival of the exotic domesticates (see “Agriculture in Native North America,” Chapter 6, p. 258). In certain other parts of what is now the United States, the earliest domesticated plants appear to have been imports from highland Mesoamerica. These plants—such as maize (corn), certain varieties of squash, and gourds—were domesticated initially in the south and then traded to the north in a down-the-line fashion, reaching the American Southwest. Yet in most of North America, the introduction (or local domestication) of these agricultural plants generally did not promote immediate or drastic shifts in dietary composition or residential mobility. In certain areas, such as the coastal Southeast, the increasing exploitation of riverine and marine life, in conjunction with hunting and gathering, led to increasing residential stability even before cultigens were incorporated into the diet. Native American lifeways also were diverse at the time of the European arrival, although no populations in the continental United States lived in urban communities as dense as those in the Basin of Mexico (discussed in Chapter 8). In the desert West, mobile hunting-and-gathering lifeways persisted until historic times. Yet in parts of the Southeast, the Midwest, the Northwest Coast, California, and the Southwest, a variety of more hierarchical political formations developed during the later precontact period. These institutions were diverse, distributed mosaically, and often short-lived in any specific locality or region. Although an increasing reliance on maize was one key factor in the development of nonegalitarian social systems in the Southeast, the Midwest, and the Southwest, such institutions were established in portions of the West Coast and in peninsular Florida in the context of sedentary hunting and gathering. Other late precontact populations incorporated the plant into their diet but remained more egalitarian and more mobile. Up to now, we have referred to prehistory as the time before the appearance of written records. In North America, the prehistoric era extends into the second millennium A.D. and ends with the arrival of populations from Europe, Asia, and Africa, who carried writing to this part of the world. The historic period begins between the sixteenth and eighteenth centuries A.D., with the specific timing varying by region, depending on the pace of European conquest and expansion. The study of the fragmentary and/or scanty historic records written by both native peoples and Europeans is called ethnohistory. In this chapter, eight North American sites are discussed (Figure 7.2). Although Mexico is geographically part of North America, the peoples who lived in
Chapter 7 Native North Americans
pri31057_ch07.pp266-319.qxd
10/1/09
8:06 AM
Page 269
most of that land area are culturally part of Mesoamerica and are discussed in Chapter 8. In time, they span the late Archaic (Poverty Point) to the era of European contact (the Draper site and Ozette). These sites—ranging from the southern Arizona desert (Snaketown) to the floodplains of the middle Mississippi (Cahokia) to the arid lands of the Colorado Plateau (Chaco Canyon)—capture some, but certainly not all, of the tremendous variability characteristic of the later prehistory of native North America. In fact, in focusing on the above sites, as well as Hopewell (Ohio) and Moundville (Alabama), we have selected rather elaborate ancient settlements, many of which have attracted archaeological attention for decades. Many of these sites are much larger and were occupied far longer than the average prehistoric North American settlement; hence, they are relatively well known. Native Americans also lived in small communities and utilized many temporary camps, but such sites are generally less well known and therefore are not featured here. Poverty Point (Louisiana) is one of the few late Archaic period sites north of Mexico where monumental earthworks were constructed. Poverty Point exemplifies certain late Archaic trends in the southeastern United States: decreasing mobility, experimentation with fired clay, and an emerging role for cultivated plants in the diet. Over the next centuries, those trends became increasingly important in several areas outside the Southeast as well. For example, roughly coincident with the decline of Poverty Point in the middle of the last millennium B.C., the Adena (ah-DEEN-ah) complex developed in the Ohio River Valley. Adena sites tend to be associated with earthen burial mounds, larger and more elaborate assemblages of grave offerings, rudimentary cultivation (of food plants as well as tobacco), and friable (soft, porous) pottery. Mobile lifeways still may have been common, but the presence of circular houses and/or ceremonial structures, as well as the increasing abundance of pottery, suggests greater residential stability at Adena sites relative to earlier occupations in the region. In many ways, the Adena complex was directly ancestral to later Hopewell sites in southern Ohio (and surrounding states). The Hopewell site, one of the richest of these settlements, was excavated on the farm of M. C. Hopewell in the early 1890s to provide artifacts for an anthropological exhibit at the 1893 World’s Columbian Exposition in Chicago. The unequal distribution of grave offerings at Hopewell sites may be evidence of social differentiation, but it is not clear how marked those distinctions were or whether they were inherited. A similar debate concerns the nature of social organization at Poverty Point. However, by the end of the first millennium A.D., on the floodplains of the Mississippi, such inherited social distinctions appear to have been well entrenched, at least at the Cahokia site in East St. Louis, Illinois. This ancient settlement, which contains the largest precontact pyramid built north of Mexico, includes ample indications of unequal access to wealth and power. Cahokia was one of the largest and most impressive Mississippian centers. Increasing reliance on corn farming at Cahokia is part of a trend found at later prehistoric sites across the eastern United States, although many Mississippian peoples continued to hunt, gather, and fish as a supplement to farming. Moundville, overlooking the Black Warrior River in Alabama, was one of several important later Mississippian centers (along with Etowah [EH-toe-wah] in Georgia). Although these two centers are well known, in part because of their monumental mound complexes and the elaborate artifacts recovered during excavations, they were not entirely typical of corn-farming peoples of the late prehistoric period who resided in what is today the eastern United States. Archaeological findings, as well as the accounts from early European explorations into the Southeast, indicate that many indigenous peoples lived in smaller, dispersed communities and were organized in less hierarchical social formations than are evident at great centers such as Moundville and Etowah.
Adena A burial mound complex that
developed in the Ohio River Valley toward the end of the last millennium B.C. Mississippian The collective name
applied to the agricultural societies that inhabited portions of the eastern United States approximately A.D. 800– 1700. Mississippian peoples constructed earthen platform mounds and shared certain basic cultural conventions.
_ Introduction The Diversity of Native American Life
269
pri31057_ch07.pp266-319.qxd
10/1/09
8:06 AM
The first Americans north of Mexico have figured as little more than a few stereotyped and contradictory images in most accounts of the New World. They are generally regarded as savages with no past and no future, sparsely inhabiting a continent, and childlike both in their generosity—“selling” Manhattan Island for $24 worth of “trinkets”—and in their ferocity—scalping the victims in battle. In the West they have been seen as a jumble of opposites, as the earlier “only good Indian is a dead Indian” attitude has given way to a latter-day Hollywood “good Indian” peacemaker juxtaposed against the “bad Indian” fighter. Only most recently can a “good” Indian fight for his people’s independence. Contemporary Indians are seen as a beaten people on reservations, sharing in a general “culture of poverty,” a curiosity for passing tourists, while the notion persists, despite census figures to the contrary, that Indians are “vanishing” as a racial and sociocultural entity. —Eleanor B. Leacock (1971)
longhouse A wooden structure that
is considerably longer than it is wide and served as a communal dwelling. Hohokam One of three major cultural
traditions of the American Southwest during late prehistoric times. The Hohokam were centered in the deserts of southern Arizona. Mogollon One of three major cultural traditions of the American Southwest during late prehistoric times. The Mogollon were centered in the mountainous areas of southeastern Arizona and southwestern New Mexico.
_ _
Ancestral Pueblo One of three major cultural traditions of the American Southwest during late prehistoric times. The Ancestral Pueblo were centered in the northern Southwest, on the high plateau of the Four Corners region.
270
Page 270
The late prehistoric groups of the Northeast and southern Canada also appear to have been organized less hierarchically. Their villages, which were composed of multifamily longhouses, lacked the monumental constructions that were found at Moundville and Cahokia. The Draper site, a large Iroquoian village in southern Ontario, was almost completely excavated and thus provides a good picture of the village plan and community organization in the Great Lakes region just before European contact. One difference between the maize-farming peoples of the late prehistoric eastern United States and those in the West is the greater reliance on water control (irrigation) in the latter area. In the dry, changeable climate of the southwestern United States, rainfall farming would have been a risky endeavor at best. This distinction is clearly indicated in the discussion of the Hohokam of the Sonoran Desert of southern Arizona. Snaketown, an early Hohokam community between Phoenix and Tucson, was for a long time the only example of an Arizona desert village. However, spurred by the contemporary construction boom in the region, additional archaeological fieldwork at other sites has supplemented and broadened our understanding of these early villagers, documenting their evolution from local hunting-and-gathering peoples. Nevertheless, Snaketown remains a key site in the region. Early-twentieth-century scholars postulated that the Hohokam migrated from Mexico; such explanations are no longer necessary to account for the transition to sedentary farming communities. The emergence of the Hohokam tradition during the first millennium A.D. occurred with the rise of two other southwestern cultural traditions, the Mogollon and the Ancestral Pueblo. Although these traditions were not completely discrete, they do help characterize the great cultural variation that marked the prehistoric southwestern United States. For the most part, the indigenous peoples of the Southwest lived in small, dispersed communities; therefore, it is not surprising to find so much diversity in diet, ceramic styles, and even residential and nonresidential construction. Periodically, the peoples of the American Southwest clustered into larger and more hierarchically organized multicommunity social networks. In general, these times of nucleation were short-lived and spatially localized. At Chaco Canyon on New Mexico’s dry Colorado Plateau, the period around A.D. 900 witnessed the beginning of perhaps the most spectacular of such episodes. Although this buildup of construction and population nucleation was relatively brief, it left an impressive record. Although the focus of this chapter is on the agricultural inhabitants of indigenous North America, we also examine the late prehistoric coastal village of Ozette in Washington. As early as 500 B.C., ranked social distinctions may have developed among certain hunter-gatherer populations in the Pacific Northwest. Although Ozette dates much later, the village’s destruction and burial by a mudslide resulted in unusually good archaeological preservation. As a result, it provides a clearer picture of Northwest Coast hunter-gatherers than can be gained from earlier sites in the area. In “Images and Ideas” (p. 317), we briefly discuss the changing EuroAmerican perspective on North American Indian sites. The archaeology of North America is extremely important because it provides a historical context for, and helps establish the rich heritage of, Native American peoples—the first inhabitants of the United States. We also raise the important issue of ethnographic analogy and discuss the less-than-perfect fit often found between ethnohistoric accounts and the archaeological record.
Chapter 7 Native North Americans
pri31057_ch07.pp266-319.qxd
10/1/09
8:06 AM
Page 271
Ozette Draper
Hopewell
Mound City
Cahokia Chaco Canyon PACIFIC
Watson Brake
OCEAN
ATLANTIC
Etowah Moundville Poverty Point
Spiro
Snaketown
OCEAN
Figure 7.2 Location of and timeline for Native American sites. Sites emphasized in text Sites mentioned in text
Poverty Point Hopewell Cahokia
site at its largest extent
Moundville Draper Snaketown Chaco Canyon
episode of occupation
Ozette 1500
1000
500
500
B.C./A.D.
1000
1500
Present
_
Year
Introduction The Diversity of Native American Life
271
pri31057_ch07.pp266-319.qxd
10/1/09
8:06 AM
Page 272
Site
:
Poverty Point Ancient earthworks in southeastern North America
Figure 7.3 New implements found at some late Archaic sites include simple, fiber-tempered pottery bowls.
The early Archaic hunter-gatherers of the southeastern United States lived in small, impermanent camps and followed a seasonal pattern of life based on a broad spectrum of plant and animal foods. The most important resources included nuts, acorns, berries, and roots and animal species such as deer, elk, bear, fox, wolf, squirrel, raccoon, opossum, beaver, otter, freshwater mussel, fish, turkey, and migratory waterfowl. By the middle of the Archaic period, around 4000 B.C., population growth contributed to reduced residential mobility and the increasing differentiation of local cultural traditions. In coastal and riverine settings, the Archaic foragers often returned to the same location year after year in scheduled seasonal rounds to exploit localized and predictable resources. For example, deep Archaic shell middens that have been excavated along the lower Tennessee River represent the accumulated refuse of small communities that occupied these same
Figure 7.4 Aerial view of Poverty Point, showing the C-shaped earthen embankments on the west bank of the Macon Bayou.
steatite Soapstone, a variety of talc
_ _
with a soapy or greasy feel, often used to make containers or carved ornaments.
272
Chapter 7 Native North Americans
sites repeatedly at the time of the year when shellfish were abundant. By 1500 B.C., before the end of the Archaic period, several plant species were cultivated in the midwestern and southeastern United States, but foraging remained the dominant subsistence strategy. One of the first crops grown, gourds were used largely as containers. Their seeds are edible, and it appears that a reliance on this part of the plant increased over time. Other plants associated with domestication included marsh elder, goosefoot (Chenopodium), and indigenous varieties of squash. New tools, including heavy ground stone implements for grinding seeds, were added to the simple portable toolkits of earlier Archaic groups. There also was an increase in food storage, and at some sites, a fairly simple and friable fiber-tempered pottery, in which fibrous plant inclusions were mixed with the clay, was made (Figure 7.3). Stone containers of carved steatite (soapstone) and sandstone are found
pri31057_ch07.pp266-319.qxd
10/1/09
Mound A
8:06 AM
Page 273
Mound B N
Geometrical earthworks Motley Mound M
ac o
n Ba
Figure 7.5 A plan of the earthworks and other structures at Poverty Point. The two largest structures, Motley Mound and Mound A, were constructed of basket-loaded dirt; some researchers believe they may have represented birds.
yo u
Floodplain 0
1 km
as well. An increasing reliance on domesticated plants (as well as other riverine resources), in conjunction with the use of less portable tools, signals a change to greater residential stability. Monumental constructions were not frequently built during the Archaic period. Out of hundreds of contemporary sites, there are only about two dozen Middle/Late Archaic settlements with small mound complexes in the lower Mississippi Valley. These sites typically have one or two conical mounds measuring between 1 m (3 ft) and 5 m (16 ft) high. The largest and most securely dated is the Watson Brake site in northeastern Louisiana. Constructed before 3000 B.C., the site’s 11 mounds, from 1 m (3 ft) to 7.5 m (25 ft) high, are connected by low earthen ridges to enclose an ovalshaped area of approximately 8 ha (20 acres). The plant and animal remains recovered at the site indicate that Watson Brake was occupied seasonally. Although no evidence of domesticated plants has been recovered, the Archaic hunter-gatherers who constructed the site were already collecting the wild ancestors of several plants that were later domesticated. Excavation has revealed few clues about the purpose of the mounds and enclosure at Watson Brake. A later, much larger complex of great earthworks, involving the movement of as much as 750,000 cu m of earth, was constructed between 1600 and 1200 B.C. at Poverty Point in northeastern Louisiana (Figure 7.4). In spite of increasing evidence of an Ar-
chaic mound-building tradition in the lower Mississippi Valley, the largescale construction at Poverty Point continues to dwarf contemporaneous Late Archaic settlements in North America. The monumental construction at Poverty Point provides additional evidence for a more sedentary residential pattern and emergence of greater social differentiation than was present elsewhere in eastern North America at that time. Poverty Point is located on the Macon Ridge overlooking the Macon Bayou and the floodplain of the Mississippi River. A portion of the site has been eroded by recent stream action; nevertheless, what remains is impressive. The main complex at Poverty Point is a set of six concentric earthen ridges that form a large semicircle, the outer one of which measures 1.2 km (0.75 mi) in diameter (Figure 7.5). The ridges are spaced approximately 45 m (150 ft) apart, and each one is composed of a series of separate smaller ridges, averaging 2 m (6 ft) high and 24 m (80 ft) wide. The presence of numerous post molds or postholes, soil stains indicating the past placement of wooden posts, as well as hearths and pits, indicates that the ridgetops were used as living surfaces. The level area encircled by earthen embankments is close to 14 ha (35 acres) and contains a central plaza that likely was used for ritual activities. To the west of these earthworks sits a large mound more than 21 m (70 ft) high and 200 m (655 ft) long. A long
post mold The circular remains, often
just a dark stain in the soil, of a wooden post that formed part of the frame of prehistoric structures. Also called a posthole.
Site Poverty Point
273
_
pri31057_ch07.pp266-319.qxd
10/1/09
8:06 AM
Page 274
Figure 7.6 Examples of small baked clay objects found at Poverty Point and elsewhere in the lower Mississippi Valley. Used for cooking in earth ovens, distinct shapes have different thermal properties, which may have permitted some control over temperatures.
equinox A time when the sun crosses
the plane of the equator, making night and day the same length all over the earth, occurring about March 21 and September 22. chert A dull-colored, subtranslucent
rock resembling flint that was often used for making flaked stone tools. galena A common heavy mineral
that is the principal ore of lead. hematite A common heavy mineral that is the principal ore of iron. slate A fine-grained rock with a dull,
dark bluish-gray color that tends to split along parallel cleavage planes, often producing thin plates or sheets. jasper A high-quality flint, often
highly colored, often used as a raw material for the manufacture of stone tools, beads, and other ornaments.
_ _
lapidary Of or related to the practice of working or cutting precious or semiprecious stones, usually for ornamental use.
274
ramp descends from the mound and faces the semicircular complex. From the mound, it is possible to view the vernal and autumnal equinoxes across the center of the earthworks; it is not clear, however, whether this alignment was intentional. A small conical mound lies about 0.5 km (0.3 mi) north of the large mound. Another large mound, similar in form to the tall ramped mound, is 2 km (1.2 mi) away. Gully erosion down the sides of the large mounds has revealed that they were built with basketloads of clay fill. The early population of Poverty Point is not well established, although the earliest excavators of the site, James Ford and C. H. Webb, estimated that several hundred houses may have been occupied. Other estimates range as high as several thousand people.
Chapter 7 Native North Americans
Smaller settlements, which lacked earthen mounds and contained fewer than 60 inhabitants, have been found up to 15 km (9 mi) away from the site. Similar settlement clusters, containing both large and small sites, also have been reported along tributary streams in other parts of the lower Mississippi Valley. Yet none of the associated earthworks is nearly as large as the impressive mounds at Poverty Point. Most sites with artifacts similar to those found at Poverty Point occur in locations where riverine resources were available. The presence of stone adzes or hoes (some of which are well worn) originally led some researchers to suggest that cultivation may have been practiced on the floodplains. Yet ethnobotanical studies have not revealed a significant presence of cultigens. Of the indigenous eastern North American domesticates, squash is most abundant. But nuts, acorns, deer, and fish seem to have constituted equal or greater portions of the diet than cultivated plants. The stone adzes (or hoes) may have been used for clearing vegetation or perhaps as digging tools for constructing the earthworks. At Poverty Point, hunting, fishing, and gathering were important means of subsistence. Very little pottery, except for some nonlocal ware, has been found. However, thousands of small baked clay objects have been recovered. Most of these fired concretions, sometimes referred to as Poverty Point objects, are smaller than a baseball and were made in a variety of forms, including balls, cylinders, and odd finger-squeezed shapes (Figure 7.6). These baked clay forms are thought to have served for cooking in earth ovens in place of rocks, which are scarce in the alluvial areas of the lower Mississippi. In this method of cooking, the clay objects were placed in pits, with fires built on top of them. Once they were heated, food was placed in the pit and roasted by the heat radiated from the clay objects. Large quantities of such ornamental objects as polished beads, pendants, and effigies were manufactured at Poverty Point, even though little stone is found on the marshy coastal plain
pri31057_ch07.pp266-319.qxd
10/1/09
8:06 AM
where the site is situated (Figure 7.7). Flints and cherts, dull-colored rocks resembling flint, were imported from the north, and steatite from the Appalachians. Other imported stone included galena (lead ore), hematite, slate, and jasper (a high-quality flint). Poverty Point was strategically located between the sources of some of these materials and the more heavily inhabited areas to the south. A well-developed lapidary (stoneworking) industry is one of the key features that distinguishes Poverty Point and surrounding sites. At some sites, distinctive microblade drills were manufactured and used to perforate stone and bone artifacts (Figure 7.8). The distribution of stone debris and finished artifacts indicates that this craft was to some degree specialized, in the sense that different communities and specific sectors within larger sites (such as Poverty Point) engaged in the manufacture of different kinds of goods. Thus the finished items as well as the raw materials probably were exchanged over great distances throughout the Southeast. Poverty Point and the specific cultural tradition associated with the site began to decline and change around 700 B.C., just as mound building increased dramatically in the Ohio Valley to the northeast. In these northern woodland areas, the construction of burial mounds implies an increasing preoccupation with death and mortuary ritual. Common practices included cremation and the placement of exotic objects with the dead. Ornaments made of marine shell, unsmelted native copper, jasper, slate, and greenstone occurred with greater frequency in these contexts. Throughout the woodlands of the eastern United States, ceramic containers became more prevalent during the last centuries B.C. The cultivation of squash and local weedy plants, which still were not staples, also became more important. These developments were foreshadowed at Poverty Point, but centuries elapsed before northeastern Louisiana again saw the scale of monumental earthen construction evident in the region at the end of the Archaic.
Page 275
Figure 7.7 Polished stone beads, pendants, and bird effigies from Poverty Point were made from a variety of stone materials that were imported to the site from many parts of the central United States.
Figure 7.8 Flint tools from Poverty Point, including microblade drills (three middle rows) that were used to perforate stone and bone artifacts. Almost 30,000 of these flint microtools have been recovered at Poverty Point alone. Site Poverty Point
_
275
pri31057_ch07.pp266-319.qxd
10/1/09
8:06 AM
Page 276
Site
:
Hopewell Prehistoric artisans and mound builders
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on Adena culture, see the Internet exercises on your online learning center.
_ _
In eastern North America, marsh elder and goosefoot may have been domesticated as early as 2000–1500 B.C., but it was not until the end of the Archaic that some prehistoric peoples in the region increasingly relied on the harvesting of these and other native plants. At archaeological sites dating before 500 B.C., the scarcity of seeds of these early domesticates suggests that these plants mainly served to buffer temporary shortages and were not yet true staples; hunting and gathering continued to provide the primary dietary resources. Nevertheless, many of these local cultigens, which produced a variety of seeds, had the advantage of being easily stored. Indigenous varieties of squash also may have been cultivated initially for their seeds rather than their flesh. These early Archaic experiments in cultivation laid the foundation for later
Figure 7.9 An artist’s reconstruction of two stages in Hopewell burial mound construction. Cremation and burial occurred in a sacred enclosure that was later covered by a mound of earth.
276
Chapter 7 Native North Americans
developments, such as Hopewell in the Midwest (100 B.C.–A.D. 400). After 500–200 B.C., native cultigens became increasingly important at many sites in eastern North America. Nevertheless, food production did not completely replace hunting-and-gathering economies. Foraging remained an important part of survival; Hopewell subsistence also included a variety of wild plants and animals, including seeds, nuts, deer, turkeys, and fish. By around A.D. 100, maize was introduced into eastern North America from the West (possibly from the American Southwest, which in turn received the plant from highland Mexico, where it was originally domesticated). Yet after its appearance, maize remained a minor crop for more than 500 years in eastern North America. About the time of the decline of Poverty Point in the lower Mississippi
pri31057_ch07.pp266-319.qxd
10/15/09
12:37 PM
Page 277
Figure 7.10 Location of Hopewell sites and other burial mound sites. Hopewell materials are found throughout the Midwest, but the focus was southern Ohio.
Lake Erie
Source: Bradley Lepper, Ohio Archaeology: An Illustrated Chronicle of Ohio Ancient American Indian Cultures. Orange Frazer Press, p. 112.
Adena
Mound City r ve
o
hi
O
Ri
Scio
to R
iver
Hopewell
Ohio
Rive
r
Valley, mound building and longdistance exchange intensified to the north in the Ohio River Valley. Elaborate earthworks—some as large as 100 m (330 ft) or more in diameter—were constructed in the shape of circles, squares, and pentagons. These earthen features appear to have been sacred enclosures rather than defensive works, perhaps serving to affirm and strengthen group cohesion and to establish the link between a local group and its social territory. During the last centuries B.C., the construction of burial mounds became an important part of mortuary activities throughout the midwestern United States (Figure 7.9). Early burial mounds contained up to three individuals in a log tomb. Cremations often were found
in simpler graves. The unequal distribution of grave goods suggests that social distinctions were marked at death in some (though not all) societies in the eastern United States. The more elaborate graves often contained greater quantities of highly crafted and exotic grave goods that possibly served as markers of rank. These items included gorgets, circular ornaments that were flat or convex on one side and concave on the other; axes, bracelets, beads, and rings made from copper imported from northern Michigan; carved tablets, some with abstract zoomorphic designs depicting birds of prey; and tubular pipes for smoking. The Hopewell tradition first appeared in Illinois around 100 B . C .
gorget A circular ornament, flat or
convex on one side and concave on the other, usually worn over the chest.
Site Hopewell
277
_
pri31057_ch07.pp266-319.qxd
10/1/09
Figure 7.11 A plan of the earthworks at the Hopewell site. Within the large rectangular embankment, or Great Enclosure, were smaller earthworks, including a circle and a D-shaped feature surrounding the largest mound at the site.
8:06 AM
Page 278
N
Spring Spring Gateway Artificial channel
Spring
Gateway
Gateway Gateway
Low bottomland
0
Hopewell Interaction Sphere A complex network involving the exchange of goods and information that connected distinct local populations in the midwestern United States from approximately 100 B.C. to A.D. 400. mica A colored or transparent mineral
silicate that readily separates into very thin sheets. effigy A representation or image of a
person or an animal. panpipe A wind instrument consist-
ing of bound sets of short pipes in graduated lengths.
_ _
celt An implement shaped like a chisel or an axe; may be made of stone or metal.
278
200 m
Material objects associated with this tradition quickly spread as far as upper Wisconsin, Louisiana, and New York, although the core of the phenomenon remained in the Midwest. The Hopewell tradition was not a single cultural group or society; rather, it was an exchange system for goods and information that connected distinct local populations. The complex trade network that defined this tradition has been referred to as the Hopewell Interaction Sphere. Goods that entered the network came from across the continent, including copper from the upper Great Lakes region and Georgia, obsidian and grizzly bear teeth from Yellowstone, chipped stone from Minnesota and North Dakota, galena from Illinois and Wisconsin, shell and shark teeth from the Gulf Coast, silver nuggets from Ontario, and mica and quartz crystals from the Appalachian Mountains. Although goods entered the Hopewell Interaction Sphere from a wide area, the focus was the Scioto River Valley in the south-central part of Ohio (Figure 7.10). Monumental mounds and extensive geometric earthworks were erected there between 100 B.C. and A.D. 400. House structures and village debris occur both within and outside the
Chapter 7 Native North Americans
k of Pai h For Nort
nt Creek
Burial mound Ditch Embankment
walls of these impressive earthworks. The Hopewell site in Ross County, Ohio, covered 45 ha (110 acres) and contained at least 40 mounds (Figure 7.11). Most of these mounds were small, dome-shaped structures, but one mound was 9 m (30 ft) high, 152 m (500 ft) long, and 55 m (180 ft) wide, and contained over 250 burials, as well as concentrations of grave offerings. Another Hopewell burial mound complex at Mound City, Ohio, has at least 24 mounds inside a large square enclosure covering 5.2 ha (13 acres). Not all the mounds at the Ohio Hopewell sites contain burials; some are effigy mounds alone. Ohio Hopewell graves contain many objects made from unsmelted native copper, such as earspools, gorgets, beads, pendants, and panpipes (primitive wind instruments). Heavy sheets of copper were used to make breastplates, and large copper nuggets were fashioned into axes, adzes, celts (Figure 7.12), and awls. Sheets of mica were cut into serpents, human hands, heads, swastikas, and bird talons (Figure 7.13). Other items found in the graves include flint tools, imported conch and other shells, teeth from alligators and sharks, turtle shells, and grizzly bear canines. A certain kind of fine pottery was cre-
pri31057_ch07.pp266-319.qxd
10/1/09
8:06 AM
ated to be used only as grave offerings. Charms were made from galena and quartz crystals. Carved stone pipes were one of the most important trade items (Figure 7.14). These pipes feature an animal effigy (beaver, frog, bird, bear, and even human) on top of a rectangular platform. The bowl of the pipe was positioned in the animal’s back, and the smoke was drawn through a hole exiting at the end of the platform. Several of the most elaborate Ohio Hopewell sites were excavated prior to 1900, before the use of modern equipment or contemporary standards of archaeological data recovery. Consequently, few of these contexts are well dated. The early excavators concentrated on the mounds where “museum pieces” could be recovered. As a result, Hopewell is known largely from these features and their contents. Less is known about Hopewell community patterns. Although many settlements were small and occupied seasonally, larger, more nucleated settlements in major river valleys were inhabited year-round. Because most of the earlier excavations were in burial mounds, some archaeologists initially thought the elaborate artifacts that were recovered represented a ritual burial cult. Yet the subsequent discovery of some of these same items in domestic settings indicates that they were more likely statusspecific objects that functioned in various civic-ceremonial contexts and eventually were deposited with the
Page 279
Figure 7.12 A copper celt. Large copper nuggets from the shores of Lake Superior were worked into a variety of forms, in addition to celts.
dead. The nature and quantities of burial goods may reflect the accomplishments and status of individuals during their lives. Only rarely were children interred with unusually large amounts of grave offerings. The Hopewell groups appear to have been organized in a number of socially differentiated polities or societies in which people reached high status principally through their individual achievements, and less through an established structure of inherited ranks. In general, it is very difficult to reconstruct the religion, ideological beliefs, or ceremonial activities of prehistoric peoples because written records are unavailable. Yet Robert L. Hall, emeritus professor at the University
Figure 7.13 A mica sheet that was fashioned into a large effigy of an eagle talon, 28 cm high.
Figure 7.14 A platform pipe in the shape of a panther. The steatite pipe is 16 cm long. Smoke was drawn through the platform below the animal’s hind legs.
_ Site Hopewell
279
pri31057_ch07.pp266-319.qxd
10/1/09
atlatl A spearthrower, or wooden
shaft, used to propel a spear or dart.
8:06 AM
of Illinois–Chicago, focusing on the symbolic importance of the famous Hopewell platform pipes, provides one potentially promising example. Drawing an analogy with postcontact customs of eastern native North Americans, Hall argues that peace pipe ceremonialism served to mediate interaction over very great distances. European explorers and traders observed that violence was absolutely forbidden among the Native Americans when the pipe was being passed. Curiously, at the time of European contact, almost all smoking pipes were in the form of weapons, such as arrows. The pipes were thus thought of as ritual weapons, and during peace pipe mediations, the participants were sometimes considered to be “fighting with words.” Hall suggests that the Hopewell platform pipes also may have been ritual weapons. These earlier pipes were made in the form of the most common Hopewell weapon, the atlatl (AT-latul), or spearthrower. (The bow and arrow were not in use in Ohio at that time.) According to Hall’s interpretation, these Hopewell effigy pipes were not simply just one of many items that were exchanged; they also were part of the mechanism of exchange itself. According to Hall, ritual “peace pipe diplomacy” served to reduce regional
_ _
280
Page 280
Chapter 7 Native North Americans
differences and to promote communication and friendly contact between the disparate groups involved in the Hopewell Interaction Sphere. The Hopewell pipes were prominent items in elaborate burials, most likely because of their importance to leaders who were involved in mediation, intergroup relations, and exchange. The Hopewell Interaction Sphere began to decline after A.D. 400. Its demise undoubtedly was related to the disruption of the trade network, but we do not know exactly what factors prompted this breakdown. One explanation is that competition in rich river bottoms between emergent slashand-burn horticulturalists, relying increasingly on maize, led to greater competition between communities. Such frictions may have caused the closing of regional boundaries. In a few cases, dispersed settlements were replaced with small nucleated villages (enclosed by ditches or walls) located in defensive positions on blufftops. The exchange of luxury goods linked the interaction sphere. Once the complex trade network was broken, either because artisans could not obtain the resources they needed or because warfare and competition disrupted alliances, the repercussions were felt throughout the regional system.
pri31057_ch07.pp266-319.qxd
10/1/09
8:06 AM
Page 281
Concept
<
The Archaeology of Exchange A perspective on ancient economies
Most—if not all—human societies engage in the exchange of goods and information, both within their own population and with other groups. Yet the volume and the mechanisms of those transactions vary greatly. Prehistorically, many communities produced most of the items they required for subsistence and survival, only occasionally exchanging an ornament or food with neighboring peoples. Other villages and societies included many specialist producers who depended on trade for their livelihood. Much ancient trade involved simple barter or reciprocity, face-to-face exchanges between known participants, such as kin or trading partners. In such small-scale exchanges, the giver assumed that a trade good or gift would be returned in the future. In the ancient Near East, pastoral herders traded mutton to sedentary farmers in exchange for grain. Other ancient trade networks were more centralized and included “nodal” individuals (like a Polynesian chief) or central institutions (such as
the Inca state) that controlled the movement, or redistribution, of certain goods. Items were collected from peripheral communities by a central authority, which then redistributed all or part of what was collected. In certain cases, redistribution primarily served to balance out environmental or economic differences between participating communities. Often, such systems allowed central authorities to accumulate large surpluses, since they would return only part of what was collected. Centrally stored goods could be amassed for military campaigns, construction projects, or the relief of local emergencies and natural disasters. Today, the U.S. federal income tax system acts as a very large and complex redistributive network. Although reciprocity and redistribution were the most common mechanisms of exchange in the past, some ancient societies were involved in market transactions. For example, as we will see in Chapter 8, the market was particularly important in prehispanic Mexico. The principles of supply and
Can researchers use the study of the archaeological record to distinguish reciprocal exchange from reciprocity? If so, how might this be accomplished?
y
Figure 7.15 A large obsidian blade from the Hopewell site.
reciprocity The exchange of goods
between known participants, involving simple barter and face-to-face exchanges. redistribution The accumulation and 5 cm
dispersal of goods through a centralized agency, individual, or institution.
Concept The Archaeology of Exchange
281
_
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 282
Figure 7.16 A copper bird effigy from the Hopewell site, Ross County, Ohio. The eye is made of freshwater pearl.
demand play a larger role in marketing than they do in either redistribution or reciprocity. In contrast to redistribution, which presumes the receipt and disbursement of goods by a central figure, marketing tends to involve more individualized transactions between buyers and sellers. Trade or exchange is best documented archaeologically by the presence of foreign or exotic goods, such as when marine shell bracelets or beads are found at inland sites many kilometers from water. However, for materials like obsidian or turquoise, determination of the source requires laboratory analysis. The point of origin of these artifacts can be identified if the specific composition of the object matches the elemental “signature” of a known source. The amount and distribution of exotic goods in archaeological contexts help determine the means and the volume of exchange. At the Hopewell site, one indication of a long-distance exchange network is the presence of ceremonial blades made of obsidian that probably came from the Yellowstone area of the Rocky Mountains (Figure 7.15). A bird effigy found at Mound City, another Ohio Hopewell site, was made with
_ _
282
Chapter 7 Native North Americans
copper from an outcrop on Isle Royale in Lake Superior (Figure 7.16). Like the obsidian blades and the copper bird, most of the items that were traded long distances to the Hopewell sites were highly crafted and rare. In addition, many of the Hopewell goods were marked with a set of distinctive designs or motifs. Archaeologists generally assume that communities that share stylistic or ceremonial conventions are more likely to have been in contact with one another than contemporary communities that have adopted entirely different artistic conventions. Thus the relatively widespread distribution of these Hopewell motifs serves, albeit weakly, as an additional indicator for the long-distance interactions that characterized this ancient exchange sphere. Although archaeologists can often document exchange, its relative volume, and its directionality, the identification of the specific mechanisms through which goods changed hands is generally more difficult.
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 283
Site
:
Cahokia The largest prehistoric community north of Mexico
After the breakup of the Hopewell trade network in the middle of the first millennium A.D., a way of life dependent on maize agriculture developed along the lower Mississippi River bottomlands. Whereas the spread of the Hopewell Interaction Sphere was largely ideological, the expansion of the Mississippian tradition probably involved some population movement into alluvial valleys with prime agricultural land, as well as the adoption of new symbols and subsistence systems by local populations. A key aspect of the Mississippian symbolic system was associated with warfare. (Another important focus was fertility.) The atlatl had been the major weapon during Hopewell times. After the decline of the Hopewell phenomenon, the bow and arrow, introduced around A.D. 500–700, became the dominant weapon in eastern North America by A.D. 1000. The lush floodplains of the Mississippi River and its tributaries were ideal for cultivating maize. Yet it was not until around A.D. 800 that maize became a major crop. Beans, another imported domesticate, and squash (local and imported varieties) also were cultivated. Yet great dietary diversity continued. Excavations at several small communities in the Mississippi Valley have revealed that maize farming was added to a diet of wild game (deer, raccoon, turkey, and seasonally available waterfowl), fish, seasonal plants (nuts, tubers, fruits, and berries), and native domesticates. Many changes in material culture that were adopted as part of the Mississippian way of life are associated with the increasing use and storage of maize. The number and size of storage pits increased significantly at emergent Mississippian sites. In addition, flint hoes were common artifacts at
Cahokia and many other Mississippian sites. Population increased rapidly in some riverine areas, reflecting the more sedentary lifeways that had been adopted, the role of storable foods in easing people through lean seasons of the year, and/or changes in labor strategies that may have given selective advantages to larger households that could clear and cultivate more land. The latter would increase the economic value of children, thereby leading to larger households and regional population increase. Cahokia (kah-HO-kee-uh), the largest Mississippian center, was established on the east bank of the Mississippi, in Illinois, across the river from St. Louis. This area, just south of the confluence of the Illinois, Missouri, and Mississippi rivers, is known as the American Bottom. The soils are extremely fertile, and there is a great diversity of environmental zones, including swamps, ponds, forests, and wet prairie grasslands. The major rivers provided avenues for communication and transport that enabled Cahokia to become the hub of an extensive exchange system (Figure 7.17). Black chert came from the Ozarks, sheet mica possibly from North Carolina, native copper from Lake Superior, lead from northern Illinois, and marine shells from the Gulf Coast. Based on the anomalous quantity of stone scrapers found in domestic contexts, James B. Stoltman, emeritus professor at the University of Wisconsin–Madison, has suggested that deer hides or dried meat may have been traded south from southwestern Wisconsin. The name Cahokia was derived from an Indian group that lived in the area at the time of French colonization during the late seventeenth and early eighteenth centuries. However, the first intensive archaeological investigation
Figure 7.17 Cahokia-style red stone figure found at the Spiro site in Oklahoma. Red flint clay pipes and figures have been found at other sites in the Mississippi and Arkansas River basins. Based on compositional analyses by Thomas Emerson, of the University of Illinois, and his colleagues, these objects were crafted using stone materials from sources near Cahokia.
_ Site Cahokia
283
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 284
N To Mounds 1 and 2 Kunnemann Tract
Creek Bottom Plaza
Monks Mound
East Plazas Cahokia Cree k
To South Mound Group
West Plaza Grand Plaza
Mound 72 Woodhenge To Powell Mound Group
Figure 7.18 A plan of Cahokia, showing the dispersal of mounds and plazas across the central area of the site. The mounds occur in a variety of shapes and sizes, reflecting their different functions.
UNESCO World Heritage Site A property considered by the World Heritage Committee to have outstanding universal value. UNESCO (United Nations Educational, Scientific and Cultural Organization) encourages the protection and preservation of cultural and natural heritage that is considered to be of outstanding value to humanity. palisade A fence of posts or stakes
erected around a settlement for defensive purposes.
_ _
284
of the site was not carried out until the 1920s, by Warren Moorehead under the auspices of the Illinois State Museum and the University of Illinois. This work confirmed that the mounds were humanmade, rather than just erosional remnants. Also in the 1920s, the first aerial photographs of Cahokia, and probably some of the earliest shots of an archaeological site in the United States, were taken. Today, Cahokia is designated a UNESCO World Heritage Site, and the central part is protected, but each year the expansion of subdivisions and highways continues to destroy outlying portions of this ancient settlement. During the past few decades, many of the archaeological projects at Cahokia have been salvage operations designed to recover as much information as possible before the destruction of a portion of the site. Fully agricultural inhabitants first settled in the Cahokia area between A.D. 600 and 800. Small hamlets and villages consisted of small houses constructed in shallow rectangular basins. Several distinct villages were dispersed on the terrain that later was to become Cahokia. By A.D. 900, the population in the region had expanded, and a hierarchy of settlements had emerged. The
Chapter 7 Native North Americans
larger communities or towns contained 100–150 people. These sites often included a series of flat-topped platform mounds that served as the foundations for temples, other public structures, or elite residences. Frequently, the platform mounds were arranged around rectangular open plazas. Not long after, sites often were fortified with defensive palisades, barricades that enclosed the plazas and other nonresidential features. Houses generally were erected both inside and outside the palisades. Many of the towns were planned settlements, with rectangular single-family houses arranged in an orderly manner around the central plaza. At its peak between A.D. 1050 and 1250, Cahokia was the largest prehistoric civic-ceremonial center north of Mexico (Figure 7.18). The site encompasses more than 100 earthen mounds in an area of approximately 13 sq km (5 sq mi) and may have had a population as high as 10,000 to 20,000, although population estimates for the site vary widely. The central part of Cahokia consists of a large plaza surrounded by the giant Monks Mound and 16 other earthen platforms (Figure 7.19). This entire “downtown” area covered more than 80 ha (200 acres)
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
and was surrounded by an elaborate and massive wall. The wall served defensive purposes, since it included screened entrances and bastions. Yet it also may have been used to limit access to the central portion of the site. The people residing within the palisaded area undoubtedly were the highest-ranked individuals in the community. People of lesser status had their own houses, public buildings, and burial mound areas outside this central portion of the site. Monks Mound, the largest prehistoric structure in the United States, consisted of four platforms, with a large public structure and some related smaller structures and walls located on the summit (see “Monumental Architecture,” p. 287). The other mounds at the site occur in a variety of shapes. Conical mounds were probably used as burial mounds. Linear “ridgetop marker” mounds situated at the edges of Cahokia may have served to delineate the limits of the site. The most common mounds are large
Page 285
square platforms, several of which had two levels, such as Mound 42, a rectangular platform mound with a smaller mound on one corner. Excavations have revealed the presence of wood and postholes on top of several platforms, which are assumed to have served as sites for ceremonial structures. Also within the site area are features called woodhenges, demarcated by large upright timbers that once had been arranged in a circle. By sighting along certain marker posts, one can observe the annual sequence of the solstices and the equinoxes; woodhenge features may have served as solar observatories. Excavations by Melvin Fowler, emeritus professor at the University of Wisconsin–Milwaukee, at Cahokia’s Mound 72 have uncovered a series of burials that date between A.D. 1050 and the early 1100s. In all, there were six burial episodes, involving at least 261 individuals (118 of whom are thought to have been retainer sacrifices), and each episode resulted in the expansion
Figure 7.19 An artist’s reconstruction of Monks Mound and the central, walled part of Cahokia at its peak, around A.D. 1150.
What indicators might archaeologists use to find evidence of past warfare?
y
woodhenge A circular feature de-
marcated by large upright timbers, probably used by prehistoric groups as astronomical observatories. solstice The time of year when the
sun is at its greatest distance from the equator, occurring about June 21 and December 22. cache A collection of artifacts, often
buried or associated with constructed features, that has been deliberately stored for future use.
_ Site Cahokia
285
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Figure 7.20 The skeleton of a man lying on a burial platform made of thousands of drilled shell beads. The burial is from Mound 72.
_ _
Figure 7.21 A Cahokia pottery vessel decorated with a sophisticated spiral scroll design. This tall vase stands 23 cm high.
286
Page 286
of the mound as prior burials were covered over. During one of the first episodes, a cache of offerings, including pottery vessels, projectile points, and shell beads, was placed in a pit. In a subsequent episode, a pit containing more than 50 young women and a platform with 4 young men, who had been beheaded and behanded, were added. The closeness in age of the women argues against their having died from disease or some common disaster. Along with the 4 men, they appear to have been sacrificial victims who were dispatched to accompany powerful chiefs after death. One burial contained a possible chief who was laid out on a litter composed of thousands of drilled shell beads and surrounded by offerings (Figure 7.20). Near this figure were 6 individuals with elaborate grave goods, including rolled sheet copper, mica slabs, polished stones, and caches of arrowheads. The complexity of the Mound 72 burial sequence, the sacrifices, and the grave goods document elaborate burial ceremonialism and marked social differentiation at Cahokia.
Chapter 7 Native North Americans
Cahokia was a great center that grew rapidly to dominate the floodplain for miles around. At its height, it contained a disproportionate amount of elaborate public architecture compared with the settlements in its surrounding hinterland. The artisanship and diversity of the goods present in burial and domestic contexts at Cahokia were unmatched at the smaller settlements (Figure 7.21). In addition, there is every reason to believe that Cahokia was well planned and that its construction required the control of a reasonably large, organized labor force. Cahokia appears to have been at the pinnacle of a hierarchically organized and complex social system that was centered on the rich American Bottom for several centuries. After A.D. 1250, Cahokia continued to be occupied, but many of the site’s mounds fell into disuse. Population at the site and throughout the American Bottom was greatly diminished as people moved into surrounding uplands. The decline of Cahokia and some nearby settlements may have been related to the growing importance of other large Mississippian centers, such as Moundville in Alabama, that redirected channels of exchange, alliance, and communication. At the same time, the great buildup at Cahokia could have precipitated degradation of the local environment. Whatever led to the transition, the period after A.D. 1250 was dominated by a series of important Mississippian centers that were dispersed across the southeastern United States. Archaeological findings suggest that each of these sites was smaller and less monumental than Cahokia had been. The archaeological record on this point dovetails with early documentary accounts from the period of first white contact. The earliest European travelers in the southeastern United States encountered a number of chiefs, temple mounds, and hierarchically organized societies, but none of the reported centers matched the size or monumentality of Cahokia.
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 287
Concept
<
Monumental Architecture Monks Mound at Cahokia
Monumental construction is characteristic of complex societies. Some of the most famous structures of the ancient world are the Egyptian pyramids and the massive architecture and pyramids along the Street of the Dead in Teotihuacan, Mexico. Although large-scale constructions were erected by most ancient civilizations, the specific activities associated with these massive edifices vary from structure to structure and region to region. Some largescale architecture served primarily to commemorate the dead, and other structures served as platforms for temples or high-status residences. In certain cases, massive architectural features (elaborate tombs or gigantic palaces) were built to confirm and glorify the power of a specific ruling individual or family, but in other places, great public structures or plazas were built to hold more corporate political and ritual functions in which a larger segment of society participated. Of course, many monumental buildings served multiple functions. The study of ancient architecture allows archaeologists to identify the political and ceremonial activities carried out by peoples in the past. In some cases, the examination of ancient buildings has enabled archaeologists to estimate the size of the labor force that was required to construct them. The largest prehistoric structure in North America (north of Mexico) is Monks Mound at Cahokia (Figure 7.22). This immense earthen mound, located at the center of Cahokia, was 30 m (or 98 ft) high, as tall as a 10-story building, and larger than a city block (316 m long ⫻ 241 m wide, or 1036 ⫻ 790 ft). Covering 6.5 ha (16 acres), it contained about 731,000 cu m (25,810,000 cu ft) of earth, or the equiv-
alent of seven modern oil tankers. Monks Mound was erected in as many as 14 stages, beginning early in the tenth century A.D. The last level, the summit on which large public buildings were built, dates to the middle of the twelfth century. In 1809, a land claim of 160 ha (400 acres), including Monks Mound, was assigned to Nicholas Jarrot, who donated this area to a group of Trappist monks. The monks, who established a monastery, gardened on the first terrace of the largest mound and built their settlement on a small mound nearby. They abandoned the site only 4 years later because of disease and hardship, but it is from this brief occupation that the largest mound at the site is now known as Monks Mound. In 1831, Amos Hill moved onto the property and built his house on the summit of Monks Mound. The area was under private ownership until 1923, when it was turned into a state park. The form of Monks Mound is unique; it consists of four terraces or platforms. The lowest, or first, terrace extends across the south end of the structure and stands approximately 11 m (36 ft) above the surrounding ground surface. A projection that appears to have been a ramp leading to the ground is situated near the center of the southern end of the terrace. The southwest corner of this terrace is slightly higher than the rest. Excavations have shown that large public buildings were erected in this area; it likely was an important focal point for the Cahokia community around A.D. 1100. These public buildings were later covered over by another platform mound that also had a public structure on top of it. This platform mound was rebuilt several times. Later, a ridge was
As a product of social effort, patterns of public architecture may communicate the nature and scale of social order. —Jerry Moore (1996)
_ Concept Monumental Architecture
287
pri31057_ch07.pp266-319.qxd
10/1/09
Figure 7.22 Monks Mound at Cahokia as it appeared in 1892, looking north. The base is approximately 300 m north to south, more than 210 m east to west, and 30 m above the surrounding fields at the highest point.
8:07 AM
constructed connecting this platform with the main mass of Monks Mound to the northeast. Around A.D. 1700, Native Americans returned to the site and buried several individuals in this part of the mound. The second terrace, extending from the first terrace back toward the northwest corner of the mound, actually consists of two flat-topped platforms at an elevation of 19 m (62 ft). These platforms were built sometime after A.D. 1250. The third and fourth terraces, situated in the northeast part of the structure, form the highest points of the mound. The uniqueness and size of Monks Mound attest to its function as a symbol of civic-ceremonial power. Yet to many archaeologists, Monks Mound is more than a sacred place where annual ceremonies probably took place. The identification of ramps and stairways on the mound suggests that access to
_ _
288
Page 288
Chapter 7 Native North Americans
the apex of the mound, which stood far above the surrounding floodplain, was restricted, possibly limited to a small segment of Cahokia’s population. Its monumentality provides an indication of the large amount of labor needed for its construction. Yet how many households were involved in this work effort and how much labor each domestic unit contributed remains an issue of debate. Finally, Monks Mound, still visible today for miles around, also was the focal point of a powerful polity, a place where Cahokia’s rulers surely must have displayed their political influence for many to witness. In fact, Cahokia remained the largest population center ever to exist north of the Rio Grande until the eighteenth century.
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 289
Site
:
Moundville A late Mississippian center in Alabama
Major changes in subsistence, material culture, and settlement patterns took place in the southeastern United States after A.D. 700. As with groups living on the Mississippi River bottomlands, corn became more important in the diet. Shell-tempered pottery appeared— a technological breakthrough that made possible the construction of larger, more durable ceramic vessels. Although most settlements in the Southeast remained small, a few larger communities with several small pyramidal earthen mounds arranged around an open plaza were established. While none of these sites yet approached the size and complexity of Cahokia, the presence of residences on the tops of some of the mounds, which were continuously rebuilt, enlarged, and inhabited for generations, suggests that a more stable or institutionalized form of elite status (perhaps inherited leadership positions) had developed in portions of the Southeast. Although the shift to maize did not occur simultaneously throughout the area, large quantities of maize were grown in much of the Southeast by A . D . 1200. Wild plants and animals continued to be important sources of food, but maize agriculture (supplemented by beans and squash) was the economic foundation of these complex societies. The social and political hierarchies of this period were manifested in public architecture, as civicceremonial centers proliferated across the region. One of the largest of these centers was Moundville, a Mississippian community located on a bluff overlooking the Black Warrior River in west-central Alabama (Figure 7.23). Most Mississippian settlements were linked by political, economic, and social ties into larger regional polities (political organizations) that varied greatly in size and complexity.
Some were small and simple, each consisting of a single center and its immediate hinterland. Others were much larger, consisting of major centers, minor centers, and villages. The larger polities, which may have had several levels of chiefs, did not emerge until after A.D. 1200. Many of the groups actually consisted of semiautonomous polities linked together by alliances. Such formations, or confederations, were constantly subject to fragmentation and realignment, especially as distance from the paramount center increased. Relationships between communities also were maintained through exchange. Such materials as copper, marine shell, mica, galena, fluorite, and bauxite were moved over great distances in both raw and finished form. At Moundville, nonlocal materials were abundant, including marine shell from the Florida Gulf Coast, copper from the Great Lakes, pottery from many areas of the Southeast, galena from Missouri, and finished ceremonial objects from Tennessee and the Spiro site in Oklahoma. Most of the artifacts made from these exotic materials tend to be associated with rich burials at Moundville, suggesting they were traded through elite channels. More domestic items, such as salt and chipped stone, probably were traded through reciprocal transactions at the household level. One of the more striking features at Moundville and other large Mississippian centers is the presence of an art style known broadly as the Southeastern Ceremonial Complex, also called the Southern Cult (Figure 7.24). It was not really a cult but a network of interaction, exchange, and shared information that crossed regional and local boundaries. Yet the specific iconographic motifs that were used varied
The European colonization and conquest was a disaster to Indian societies. Among the most effective introductions were epidemic diseases, which reached the interior much faster than direct intervention. —James B. Griffin (1967)
Southeastern Ceremonial Complex
A network of interaction, exchange, and shared information present over much of the southeastern (and parts of the midwestern) United States from around A.D. 1200 until the early 1500s; also previously referred to as the Southern Cult.
Site Moundville
289
_
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 290
La k
Lake Michig an
on ur eH
Figure 7.23 Mississippian sites were concentrated on the wide floodplains of the Mississippi River and its major tributaries. Although there were thousands of Mississippian mounds in the southeastern United States, few of them remain today.
er
Cahokia
Ri
v
Ri
ver
O hio
Mi s s
iss
ip
pi
Spiro
Etowah Moundville
ATLANTIC OCEAN
Gulf o f M exi co
Figure 7.24 A black pottery vessel with bird effigy found at Moundville, 16 cm high. Birds and serpents figured prominently in the iconography of the Southeastern Ceremonial Complex.
motif A recurring thematic design element in an art style.
_ _
charnel house A house in which the
bodies of the dead are placed.
290
across space and over time. Items belonging to this complex have been found from Mississippi to Minnesota, and from Oklahoma to the Atlantic Coast, although they are most abundant at certain sites in the Southeast. Such motifs as human hands with an eye in the palm, sunbursts, weeping eyes, and human skull-and-bones are elements of this style (Figure 7.25). They appear on polished black shell-tempered pottery, are embossed on pendants made from Lake Superior copper, and are incised on imported Gulf conch shells. The most famous Mississippian cult objects are the so-called effigy jars that are decorated with human faces, some with signs of face painting or tattoos. Others represent
Chapter 7 Native North Americans
sacrificial victims, with eyes closed and mouths sewn shut. Often the effigies are shown weeping, possibly denoting a connection between tears, rain, and water in Mississippian cosmology. Some of the common motifs—wind, fire, sun, and human sacrifice—seem to share certain thematic elements with Mesoamerica, although the basis for this similarity has not been established. The greatest concentrations of Southeastern Ceremonial Complex objects occur in temple mounds at some of the major late Mississippian sites, suggesting that these goods had political and symbolic importance. Because of its large size, Moundville has long attracted public atten-
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
tion. In 1840, Thomas Maxwell, a local planter and merchant, excavated in one mound, noted the stylistic similarities with Mesoamerica, and concluded that the site was an outpost of the Aztec empire. Several small-scale investigations of the site were made by the Smithsonian Institution during the latter part of the century, but the first large-scale excavations were not carried out until 1905–1906 by Clarence B. Moore. Excavating both the platform mounds and village areas, Moore uncovered over 800 burials, many accompanied by pottery vessels and other artifacts of shell, copper, and stone. The second major episode of excavation was carried out by the Alabama Museum of Natural History. From 1929 through 1941, 4.5 ha (11 acres) of the site surface were opened, and the excavations yielded over 2000 burials, 75 structures, and many other finds. More recent work has concentrated on understanding the chronological sequence at the site. In all, excavations at Moundville have yielded more than 3000 burials and over 1 million artifacts. Based on recent studies, the history of Moundville is divisible into three episodes. Moundville was occupied initially around A.D. 1050. At first, the site contained only two small mounds and was one of several small ceremonial centers in the region. The burials from this period indicate some signs of marked, possibly inherited, social differentiation, though not to the extent found for later occupations. As noted above, this initial episode of growth was associated with a rapid intensification of maize farming. By A.D. 1200, a marked period of growth and political centralization had occurred at a time when Cahokia was already in decline. Moundville grew to over 75 ha (185 acres), and at least 30 pyramidal mounds were built around a large 32-ha (79-acre) plaza (Figure 7.26). Around A.D. 1300 the character of Moundville shifted again. The overall population declined, yet the site was still occupied by a higher-status segment of the population. Moundville continued to be an important political
Page 291
and religious center and remained a locus for mortuary activities. The mounds at Moundville were large, flat-topped earthen structures constructed to elevate temples or the dwellings of important individuals above the surrounding landscape. The mounds were built in stages, probably as part of community rituals. The two major mounds, the largest of which was 17.3 m (58 ft) high and covered more than 0.8 ha (2 acres) at its base, were located within the plaza at the center of the site along the north-south axis. The 18 earthworks surrounding the plaza are arranged into pairs of one large and one small mound. The small mounds usually include burials; the larger earthen structures do not (Figure 7.27). In the northeast corner of the site, the dwellings were larger and more complex than in other parts of Moundville. Broken artifacts that correspond to items found in the higher-status burials also are found in this probable elite residential area. Charnel houses and a sweat house are located along the margins of the plaza. Commoner residential areas were placed at greater distances from the eastern, southern, and western sides of the plaza. At its height, Moundville is estimated to have been occupied by a few thousand people.
Figure 7.25 An engraved sandstone palette from Moundville, with the eye-in-palm motif circled by two horned rattlesnakes (diameter about 31.9 cm).
Figure 7.26 An artist’s rendering of Moundville as it may have appeared at its height of occupation.
_ Site Moundville
291
pri31057_ch07.pp266-319.qxd
B la
ck
10/1/09
W
io arr
8:07 AM
Page 292
i v er rR
Plaza
N Mound Borrow pit Plaza
Palisade 0
Figure 7.27 A plan of Moundville. The spatial arrangement of earth mounds and plazas gives the impression of symmetry and planning.
iconography The study of artistic representations or icons that usually have religious or ceremonial significance.
_ _
conquistador (Spanish) A conqueror.
292
200
400 m
Several craftworking areas have been identified away from the residential zones at Moundville. One contained a large quantity of finished shell beads, unworked shell, and beadworking tools, and another yielded hundreds of large bone awls and the stones used to sharpen them. The bone awls may have been used for processing hides. Large fire hearths, caches of shell (for temper), clay, and other items indicative of pottery production were found in a third area of craftworking. After A.D. 1200, social differentiation was clearly expressed in mortuary goods. The high-status segment of the society, which totaled approximately 5% of the population, was always buried in or near mounds. These individuals were always accompanied by rare and distinctive artifacts, including copper axes and gorgets, stone disks, items belonging to the Southeastern Ceremonial Complex, and many shell beads. Each mound also contained lower-status individuals, who were accompanied only by a few ceramic vessels. The less elaborate burials could
Chapter 7 Native North Americans
have been retainers who were positioned to accompany those of higher status. The largest proportion of Moundville’s inhabitants were interred away from the civic-ceremonial part of the site. These graves generally included no more than a few ceramic vessels. The inclusion of people of both sexes and all ages in each class of burials suggests that access to goods (at least at death) was determined partially through birthright. Although social status at Moundville appears to have been inherited, warfare may have provided one mechanism for certain individuals to enhance their status. Evidence of warfare is present in the archaeological record at Moundville, as well as elsewhere in the Southeast. Many communities, both large and small, were surrounded by fortification walls and ditches. In addition, skeletal studies have revealed numerous indications of scalping, the taking of trophy heads, and the burial of headless bodies. Warfare is also prominent in the iconography of the Southeastern Ceremonial Complex. One grave at Moundville contained 11 decapitated skeletons. This grave also included ceremonial flints and shell objects engraved with elaborately dressed individuals holding severed heads. Mississippian culture persisted in parts of the Southeast until the sixteenth century. In the 1540s, Spanish conquistador Hernando de Soto encountered indigenous Creek chiefs who still lived in fortified towns with temple mounds and plazas. With the European expeditions came smallpox and other diseases against which the indigenous peoples had no immunity. These communicable diseases quickly decimated the Native American population, putting an end to many facets of a way of life that had endured for centuries.
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 293
Concept
<
Grave Offerings Indicators of social differences
Grave offerings, the items buried with individuals at death, are an important source of information about the social organization of prehistoric groups. With other aspects of mortuary ritual, such as location of the burials and the “elaborateness” of tombs, grave offerings inform archaeologists about the relative social position of the interred individuals. A person’s status during life is often reflected at death: elaborate burials and grave goods for people of high status, but few or no trappings for people of low status. Complex societies with marked social differentiation usually have a greater degree of mortuary variation than less hierarchical societies. The distribution of grave offerings relative to age and sex also may indicate whether an individual had achieved status, earned through personal accomplishments, or ascribed status, inherited at birth. Some archaeologists have argued that graves of infants or children with an unusually rich array of offerings are indicative of ascribed status, since these individuals would not have been able to achieve high status on their own. Of course, the specific manifestations of these patterns vary somewhat from one context to another. In the Mississippian societies of the southeastern United States, social differentiation was expressed in mortuary ritual. Commoners were interred in simple graves, usually grouped in communal cemeteries or scattered near dwellings. Typical grave offerings included simple shell ornaments, ceramic vessels, and domestic tools. In contrast, high-ranking people were buried in or near public buildings, often on mounds, and accompanied by elaborate offerings. The more than 3000 burials excavated at Moundville can be divided into several groups (Table 7.1). Seven
individuals had very high status; they were probably chiefs. All male, they were buried in large mounds and were accompanied by lavish grave goods, such as copper axes, copper-covered shell beads, and pearl beads. Sacrificial victims accompanied these burials (group A in the table). A second group of elaborate burials (group B) included both children and male adults; these graves also were placed in or near the mounds. This group was interred with copper earspools, stone disks, beartooth pendants, oblong copper gorgets, and artifacts decorated with symbols associated with the Southeastern Ceremonial Complex. Other groups of high-status interments (groups C and D) included individuals of both sexes and all ages, buried in cemeteries near the mounds, accompanied by shell beads, oblong copper gorgets, and galena clubs. Since rare and exotic items were buried with individuals of both sexes and all ages, it appears that status at Moundville must have been ascribed rather than achieved. Less elaborate burials at Moundville also were divisible into several groups (E–H), including a large number, mostly children and infants, that lacked any grave offerings at all. Less ornate grave goods were distributed very differently from the offerings associated with high-status burials. In the lower-status graves, the distribution of goods corresponded more closely with variation in age and sex. Graves of older adults generally contained pottery vessels, bone awls, flint projectile points, and stone pipes. Such other items as unworked deer bones, bird claws, and turtle bones were found exclusively with adults; children and infants were sometimes accompanied by toy vessels and clay toys. Stone ceremonial celts (axe heads) were found only with adult males, whereas effigy
achieved status Social status and prestige attributed to an individual according to achievements or skills rather than inherited social position. ascribed status Social status and
prestige attributed to an individual at birth, regardless of ability or accomplishments.
Graves and deposits containing human remains are important sources for archaeologists. The bones and other surviving tissues can be analyzed to determine the age and sex of the deceased and sometimes diet, diseases, and cause of death. These methods of osteology and palaeopathology are combined with mortuary analysis to understand the organization of past societies. —Michael Parker Pearson (1996) Concept Grave Offerings
293
_
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 294
TABLE 7.1 The Complexity of Mortuary Ritual at Moundville
Broad Groupings
Burial Group
Higheststatus graves
Other high-status graves
Lowerstatus graves
Number of Individuals
Infant Child % Age Adults
Characteristic Grave Offerings
Burial Context
Copper axes
Central mound
I C A
? Male
Sex
A
7
B
110
Copper earspools, bear teeth, stone disks, red or white paint, shell beads, oblong copper gorgets, galena
Mounds and cemeteries near mounds
I C A
Male and female
C
211
Effigy vessels, shell gorgets, animal bone, freshwater shells
Cemeteries near mounds
I C A
Male and female
D
50
Discoidals, bone awls, projectile points
Cemeteries near mounds
I C A
Male and female
E
125
Bowls and/or jars
Cemeteries near mounds and in village areas
I C A
No data
F
146
Water bottles
Cemeteries near mounds and in village areas
I C A
Male and female
G
70
Sherds
Village areas
I C A
Male
H
1256
No grave goods
Retainers in mounds; isolated skulls with public buildings; cemeteries near mounds and in village areas
I C A
Male and female
Source: Adapted from Peebles and Kus, 1977.
y
Although grave offerings and burial treatment can provide indicators of social differences, there also can be confounding factors and considerations. What are these potential complications, and how do they affect archaeological interpretation?
_ _
294
vessels were associated with adults of both sexes. The different grave offerings indicate that status for commoners was determined according to sex and age. Although high or low status in general was a result of birth, ranking within each group seems to have depended partly on individual achievement or role. Similar burial patterns are found at other contemporaneous sites in the vicinity of Moundville. But there is one major difference: The most elaborate burials in the mounds of the minor
Chapter 7 Native North Americans
ceremonial centers (those sites with only a single platform mound) did not contain copper axes—the badge of office—that accompanied the highestranking burials at Moundville. The highest-ranking individuals in the social system apparently resided at the region’s largest and most important ceremonial center.
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 295
Site
:
The Draper Site A late prehistoric Iroquoian village in southern Ontario
When early European explorers and fur traders arrived in the northeastern United States and southeastern Canada, the native peoples they encountered differed in several important ways from the aboriginal populations farther south. Although maize agriculture was prevalent among most of the peoples east of the Mississippi River, the native peoples of the Northeast differed from those of the Southeast in village structure and social organization. Generally, the northeastern groups were organized less hierarchically than the societies of the Southeast. In the northeastern Iroquoian communities, leaders were chosen largely for their skills in settling disputes and organizing military expeditions and for their ceremonial knowledge. Leadership roles were based more on achievement than on ascription; the latter pattern was more common among the southeastern groups. Also, northern villages were composed of clusters of longhouses, often surrounded by palisades and lacking the platform mounds found at sites in the Midwest and Southeast, such as Cahokia and Moundville. The historic Iroquoian tribal groups of the Northeast were the descendants of small, mobile bands of Archaic foragers. We have little evidence of their settlement patterns; however, a small circle of post molds at one Archaic site in Vermont suggests that they built temporary shelters framed with poles and possibly covered with bark or skins. Hunting, fishing, and collecting wild plants were the primary subsistence activities. By the middle of the first millennium A.D., people in southern Ontario were living in semipermanent settlements during spring, summer, and early fall. These sites, usually positioned near good fishing locations, consisted of unfortified clusters of round or oblong
houses. Although limited amounts of squash may have been cultivated, hunting, fishing, and collecting continued to be the main subsistence activities. After A.D. 700, major changes in settlement pattern, subsistence, and social organization occurred in the Great Lakes region. Villages were repositioned from major rivers and lakes to hilltop locations that were naturally defendable, and many settlements were surrounded by palisades. Most villages remained small (around 0.6 ha, or 1.5 acres) and were composed of small, multifamily, bark longhouses. These villages were occupied primarily in winter. By A.D. 1100, maize agriculture had become a more significant part of the subsistence regime (Figure 7.28). This important dietary shift coincided with the evolution of heartier maize varieties that were better adapted to the colder northern latitudes (where the growing season was relatively short). But even toward the end of the prehistoric era in the Northeast, maize farming did not completely supplant the hunting, fishing, and gathering of wild foods. Fish remained an especially key part of the diet for many populations. After A.D. 1300, the stockaded villages were occupied on a more permanent basis. Several archaeological indicators, including an increase in the complexity of village fortifications and a growing number of traumatic injuries found on skeletal remains, point to the intensification of warfare. The historically known patterns of Iroquoian internecine warfare and blood revenge may have been established by this time. Longhouse size expanded and village size increased. Some villages were as large as 6 ha (15 acres) and included 2000 people, although the average community was closer to 2.5 ha (6.25 acres) in size. In several
Figure 7.28 Iroquois Indian basket used for washing hominy (kernels of corn). Corn became an important component of the diet after A.D. 1100.
_ Site The Draper Site
295
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 296
Figure 7.29 An artist’s reconstruction of the main palisaded village and enclosed longhouses at the Draper site, after the village had undergone its final expansion.
sweat bath A hut or other space
_ _
heated by steam that is created by pouring water over hot stones.
296
areas, settlement pattern studies suggest that these larger villages were formed through the nucleation of several smaller communities. In these larger villages, the alignment of houses was usually much more formal than in the earlier, smaller settlements. Excavations at the Draper site, located 35 km (22 mi) northeast of Toronto, have provided a wealth of information on village organization in the late prehistoric period. Most of the settlement was excavated as part of a salvage operation initiated before the construction of a new Toronto airport. An unusually large village for late prehistoric Ontario, the site is particularly important because almost the entire settlement was excavated. The high cost of modern archaeological excavation makes it very rare for more than a small portion of a site to be examined carefully. Yet in a few cases—Draper being one—the opportunity to excavate a large segment of the original village provides an unusually detailed picture of community plan, changes in settlement size over time, and houseto-house variation. The Draper site is one of more than 15 Iroquoian villages in the Duffin
Chapter 7 Native North Americans
drainage, located in the white pine– hardwood forest region of southern Ontario. The settlement was inhabited between A.D. 1450 and 1500. This relatively short occupation is not unusual for southern Ontario, where most villages were resituated every 25–50 years, usually to new locations only 3–5 km (2–3 mi) away. Many factors have been proposed to account for these frequent relocations, including infestation of the wood-and-thatch longhouses by insects, soil exhaustion or weed competition (which would have prompted the clearing of forest for new agricultural fields), depletion of wood (for construction and firewood) and game, and community realignments resulting from disputes and other social stresses. The Draper site was composed of three spatially discrete areas of occupation: the main palisaded village (Figure 7.29), a small group of seven houses located 50 m (160 ft) south of the main village, and a lone structure (Structure 42) located on a small knoll 80 m (260 ft) to the southwest. At different times in its occupation, the main village was surrounded by three or four palisade rows (each composed of wooden posts or beams). William Finlayson, who super-
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 297
Burial pit Entrance Large pit
Sweat bath
Post mold Hearth
Sweat bath
Pit feature N
Sweat bath
Sweat bath
0
5m
vised the principal excavations in 1975 and 1978, believes the settlement was planned with defensive considerations in mind. The site began as a 1.2-ha (3-acre) village of seven to nine longhouses (accommodating roughly 400 people). All the houses were built well back from the palisades. These first houses were arranged in two clusters, with the houses in each cluster laid out according to similar compass orientations. Each set of houses may have represented a distinct social unit, perhaps a lineage or clan segment. All the longhouses had a relatively similar set of internal features, including sweat baths, cooking hearths, pits (some of which were burials), storage cubicles, and benches (2–2.5 m, or 6.5–8 ft, wide), which were placed along the house walls (Figure 7.30). Each nuclear family in the longhouse is thought to have had its own cooking and sleeping area. The sweat baths were less abundant and were probably used in a more communal fashion. According to ethnohistoric accounts of the Huron, the Iroquoian tribal group considered to be the descendants of the people at the Draper site, sweat baths usually were
taken by groups of men to prevent disease. Before the bath, the stones were heated in a large fire, then removed and put in a pile in the center of the lodge. Sticks were arranged at waist height around the pile and then bent at the top, leaving enough space between the sticks and the rocks for naked men sitting with their knees raised in front of their stomach. When the men were in position, the whole bath was covered with large pieces of bark and skins to prevent air from escaping. While in the bath, the men would often sing. These sweat-bath rituals are thought to have had an important integrative function for a group of people who resided in very close quarters. In the first Draper village, one longhouse had several characteristics suggesting a special function. This large structure had one of the highest densities of wall posts and sweat baths. It also had the greatest average distance between hearths, indicating that each family group had more space than other longhouse occupants. In addition, a special hearth was placed at one end of this somewhat unusual residential structure. Finlayson suggests that this large longhouse may have been occupied by a
Figure 7.30 Excavated longhouses at the Draper site, showing the distribution of sweat baths, hearths, burial pits, and other pit features. The longhouses are clearly outlined by post molds. Other post molds also are present inside the houses.
_ Site The Draper Site
297
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 298
Figure 7.31 This effigy pipe bowl is one of thousands of clay pipe fragments that were recovered during excavations at the Draper site. Smoking pipes played an important role in Iroquoian council meetings.
community or kin-group leader or village chief. Such special structures may have been used for council meetings, community feasts, and ceremonies (Figure 7.31). The Draper village underwent five expansions, eventually reaching a maximum population of 1800–2000 people. Before abandonment, the village was 3.4 ha (8.5 acres) in size. During each expansion episode, between three and nine houses were built. The houses often were added in clusters, with the houses having similar orientations. The additions may represent new kin segments that moved into the site. Over time, an increasing amount of space was devoted to nonhouse use; with each expansion, the houses, or clusters of houses, were positioned to create new plazas. The plazas may have served as places for village ceremonies and social activities. Such integrative events may have increased in importance as the community grew. Two of the houses that were added during
_ _
298
Chapter 7 Native North Americans
these expansions are thought by Finlayson to have been leaders’ houses. Two rectangular structures that abutted the palisade also were added. These buildings have been interpreted as houses set aside for visitors. According to ethnohistoric accounts, visitors were assigned to special cabins so that they could be closely monitored; they were not allowed to wander through the community. The longhouses in the Draper site were 14.5–75.1 m (48–247 ft) long and 6.7–7.9 m (22–26 ft) wide. The narrow range of widths was probably determined by the mechanical limitations of the construction procedures and the available material (wood). In contrast, the length variation was probably related primarily to household or kingroup size. The seven houses in the south field were smaller than those in the main village. They had a low density of pits and sweat baths, suggesting a shorter occupation. The presence of only a partial palisade, possibly used as a windbreak, suggests less concern for defense. Perhaps the occupants of this southern area moved inside the larger palisaded village when sieges occurred. Structure 42, also outside the main village, appears to have been a special-purpose structure, although its specific function remains little understood. Fragments of human bone were found on the surface before its excavation, but further study has not yet revealed additional signs of a burial area. The late precontact Iroquoian villages in the Great Lakes region were not organized as hierarchically as the larger, contemporaneous Mississippian polities in the Southeast. In the latter, labor was amassed to construct large pyramid mounds at central settlements. These focal settlements differed in size and function from the many smaller villages that also constituted the settlement system. In contrast, the differences between the Draper site and surrounding settlements were not substantial. Although Draper was somewhat larger than its neighbors, the construction remains and material items at the site
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
were basically similar to those found at surrounding communities. Because of the tremendous changes that occurred in Native American lifeways as a result of direct European contact (and sometimes before, since the spread of disease and trade goods often preceded actual face-to-face relations), the early European historical accounts present a sketchy and sometimes inaccurate picture of the late precontact period. This potential disparity between archaeological and historical records is especially troublesome in areas where major settlements, such as the large Hohokam villages of the Arizona desert, were apparently abandoned sometime before European contact (see “Snaketown,” p. 300). In such cases, the native peoples that the Europeans first encountered were not necessarily the direct descendants of earlier native populations, whose communities have been found by archaeologists. In other areas, the direct sequence between the historic and prehistoric peoples is somewhat clearer, such as for the Iroquoian groups of the Northeast. In these cases, historical information can more readily be used to test and support interpretations derived from archaeological data. Nevertheless, the careful researcher must be
Page 299
prepared to recognize diversity and changes in the archaeological and historical records. For example, in the Great Lakes region, indirect contacts spread trade goods and disease vectors to most areas before direct meetings between native peoples and Europeans took place. In fact, some scholars have argued that the large-scale Iroquoian alliances and confederacies, which at times joined up to 25 villages and are well documented in the early historical accounts, may have been formed in response to contact period processes. Such large linkages appear not to have been in place prehistorically. New trade demands and land pressures may have prompted the formation of new political structures. Warfare and trade certainly were important prehistorically in the Great Lakes region, but after European contact, they took on new forms. Scholars, such as those studying the Iroquoian peoples, who can use and compare archaeological and historical sources as partially independent records, stand the best chance of unraveling the multifaceted and often destructive processes that surrounded European arrival in the Americas and the effects of these processes on the native peoples.
_ Site The Draper Site
299
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 300
Site
:
Snaketown A desert village in the American Southwest
Figure 7.32 Mano resting on a wellused metate.
metate The stone basin, often trough-shaped, or lower part of a stone-milling assembly for grinding maize or other foods. mano The handheld part of a stonemilling assembly for grinding maize or other foods.
_ _
wattle and daub A building technique that uses a framework of poles, interspersed with smaller poles and twigs; the wooden frame is plastered with mud or a mud mixture. This building technique also was employed in the Southeast and other parts of the world.
300
Despite its often arid and unforgiving landscape, the American Southwest has a rich archaeological past. The region was first populated around 9000 B.C. by hunter-gatherers who hunted big-game species, such as mammoth and giant bison. These early inhabitants had highly mobile foodcollecting strategies and manufactured sophisticated tools for hunting and butchering game and for processing hides, wood, and bone. Their remains occur primarily as small, ephemeral kill and butchering sites characterized by animal bones and flaked stone implements. By 5500 B.C., big-game hunting had declined in importance, and emphasis shifted to smaller animals and a variety of local plant resources. In southeastern Arizona and southwestern New Mexico, these wild foods included yucca, cactus leaves and fruits, and sunflower seeds. After 2100 B.C., maize was introduced into the area from Mexico. Along with other early southwestern cultigens (beans, squash, and bottle gourds), maize had a long history of cultivation in Mexico before its appearance in the Southwest. Eventually, maize became the most important cultivated crop in the Southwest. However, for 1000 years following its introduction, most peoples of the Southwest relied surprisingly little on maize or any other exotic domesticate, and many groups retained a primarily hunting-and-gathering subsistence regime. The earliest villages in the Southwest appeared in the desert river valleys of southern Arizona during the Late Archaic period (middle of the last millennium B.C.). This more sedentary existence was marked by circular foundation pits, storage facilities, crude plainware pottery jars and bowls, jewelry made from marine shells, and maize farming. By the end
Chapter 7 Native North Americans
of the Late Archaic, agricultural villages dotted the landscape throughout the American Southwest, coexisting in some areas with groups who continued more traditional foraging practices and nomadic lifeways. In the more northerly, cooler parts of the Southwest, farming was relatively risky. Perhaps that is part of the reason many early villages were positioned defensively and generally included storage facilities. In the hot Sonoran Desert of southern Arizona, crop production was more secure (as long as sufficient water was available), villages were situated more in the open, and no specialized storerooms were constructed. The presence of basketry containers, along with metates, grinding stones, and manos, companion handstones—for grinding corn or seeds—is indicative of a more settled existence in that area (Figure 7.32). The early villages of the Southwest shared certain general characteristics. The earliest permanent dwellings were pithouses, structures in which the lower parts of the walls were actually the earthen sides of a shallow pit (Figure 7.33). The top part of the walls consisted of a framework of poles, interlaced with small twigs and then completely covered with mud on the exterior, a construction technique known as wattle and daub. Settlements often contained only two or three pithouses. In contrast, larger villages usually had one or more community, or special-function, structures in addition to pithouses. Although the earliest ceramics in the Southwest were plainwares (without decoration), distinct regional variations soon developed. During the early centuries of the first millennium A.D., the three major cultural traditions of the prehistoric Southwest began to emerge: Hohokam (ho-ho-KHAM) in
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
the deserts of southern Arizona; Ancestral Pueblo on the high plateaus of the Four Corners region (northern Arizona and New Mexico, southwestern Colorado, and southeastern Utah); and Mogollon (muh-ghee-YOWN) along the Mogollon Rim (in east-central Arizona) and in the mountains of southeastern Arizona and southwestern New Mexico. Each tradition eventually developed its own pattern of settlement and land use, architecture, community organization, and craft specialization. The discussion here of Snaketown (Hohokam) and in the next section on Chaco Canyon (Ancestral Pueblo) highlights the diversity of Native American life in the prehistoric Southwest. The Hohokam lived in the lower Sonoran Desert region of southern Arizona and adjacent Chihuahua and Sonora in northern Mexico. The area is basin-and-range country, composed of numerous more or less parallel mountain ranges rising 300–1070 m (1000– 3500 ft) above the intervening basins. The desert area of southern Arizona receives less rain than the high plateaus of the Four Corners region and is intensely hot in the summer. The land supports a rich natural flora of shrubs and cacti (saguaro, barrel, cholla, and prickly pear); mesquite and other trees and shrubs grow in the washes. Yet because of the somewhat unpredictable rainfall, irrigation is necessary in many areas for reliable maize farming. After A.D. 1, a large Hohokam community called Snaketown was settled in the Phoenix Basin, a broad, low alluvial region where southern Arizona’s two major rivers, the Salt and the Gila, come together (Figure 7.34). Snaketown is situated on an upper river terrace about 1 km (0.6 mi) from the Gila River at an elevation of 360 m (1175 ft). According to the late Emil Haury, who excavated at Snaketown in 1934–1935 and again in 1964–1965, good agricultural land, a river that could be tapped for irrigation, and a high water table were important features determining Snaketown’s location. The largest of the early Hohokam pithouse villages, Snaketown may have had as many as 100 residents soon
Page 301
after its foundation. The early habitation levels included only residential pithouses; no public buildings were constructed. However, David R. Wilcox, of the Museum of Northern Arizona, has suggested that even early in its occupational history the site may have had a central plaza area. Maize was a dietary staple, although such wild plant foods as mesquite, maguey, saguaro, and cholla also were important. By A.D. 600, the number of Hohokam villages had increased markedly, and the Phoenix Basin became the most densely populated area in southern Arizona. Many villages, including Snaketown, grew. At some sites, ball courts and platform mounds were erected (see “The Mesoamerican Ballgame,” Chapter 8, p. 356). One of the most impressive of these oval Hohokam ball courts was constructed at Snaketown (Figure 7.35). Rubber balls, also found at Snaketown, suggest that certain aspects of the Hohokam ballgame may have been similar to the game played in Mexico. Hohokam platform mounds, usually low and rectangular, were made of earth and adobe and topped with caliche (calcium carbonate) or adobe plaster. Ceremonial structures may have been positioned on the tops of these mounds. Both platform mounds and ball courts were generally erected at larger Hohokam sites, such as Snaketown. These sites may have served as political-ritual centers for the populations of smaller surrounding settlements that lacked such nondomestic structures.
Figure 7.33 Two examples of pithouses from Snaketown. The roof and wall constructions are postulated, in part, on the distribution of post molds found in the houses.
ball court A prehispanic structure
that was the site of ritual ballgames. adobe A mud mixture used to make
sun-dried bricks for buildings in arid areas.
Site Snaketown
301
_
pri31057_ch07.pp266-319.qxd
10/1/09
Figure 7.34 Snaketown during excavations in 1965. The large village at its height comprised a mosaic of houses, mounds, platforms, wells, canals, and two ball courts.
_ _
302
8:07 AM
Page 302
The Hohokam cultural tradition, marked by its red-on-buff pottery, reached its greatest spatial extent after A.D. 900, spreading across much of central and southern Arizona (Figure 7.36). At about that time, Snaketown covered an area over 1 sq km (about 0.4 sq mi, the size of the Indianapolis racetrack) and may have had a population of 500–1000 people. The plan of the village became more complex as well. At the site’s core was a central plaza surrounded by an inner habitation zone. A series of mounds and two ball courts circled this area, with an outer habitation zone beyond. In all, 60 trash-filled mounds were constructed at the site, and as many as 125 pithouses may have been occupied at any one time. At Snaketown, craftspeople specialized in the production of ceramics and developed elaborate shell- and stoneworking industries. Goods from all over the Southwest and northern Mexico were traded to the site. Some items, such as the rubber for the balls,
Chapter 7 Native North Americans
may have come from even farther south. The shell was imported from the Gulf of California and was cut into bracelets, rings, pendants, beads, and ornaments decorated with designs etched with acid (Figure 7.37). Stone was fashioned into beads, earplugs, ornaments, carved paint palettes, axes with deep grooves for hafting, and vessels with animals in relief (Figure 7.38). Lizards were a common motif. Other large Hohokam villages also were present in the region, especially in the Salt River Valley to the north. Although the residents of Snaketown dug several canals, it was along the lower Salt River of the Phoenix Basin that the Hohokam constructed the largest prehistoric irrigation system north of Mexico. At its largest extent, this irrigation network consisted of 14 main canals, ranging from several kilometers up to 20 km (12.5 mi) in length, that drew water from the river, and hundreds of smaller canals that carried water to the fields (Figure 7.39). Canal irrigation required a significant labor investment,
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 303
Figure 7.35 The large oval ball court at Snaketown had a flat floor and side walls that sloped upward, forming an embankment around the court.
first to dig the canals and then to maintain and repair them. The degree and nature of coordination across this canal system remains a matter of debate. After A.D. 1150, the spatial extent of the Hohokam and their ceramic ware decreased. Snaketown was largely abandoned, and the surrounding Gila Basin was partially depopulated. Hohokam settlement then concentrated in the lower Salt River Valley, where several villages grew very large, and the irrigation network reached environmental limitations on expansion. Pithouse villages were largely replaced by rectangular adobe enclosures built aboveground, possibly for defensive purposes. Platform mounds continued to be constructed, while ball courts declined. In some instances, residential compounds, possibly associated with higher-status households, were placed on the mounds. Multistory structures were built at some sites. Although an
explanation for these changes is not well established, they do indicate that a political reorganization occurred. When the first Spanish explorers entered the Phoenix Basin, the Hohokam were gone. The irrigation systems were abandoned, and the large villages, with their associated mounds and ball courts, lay in ruins. The Pima and Papago Indians, who consider the area to be their historic homeland, were the desert’s only inhabitants and resided in small, scattered villages. Hohokam is their name for the people responsible for the ruins; the word can be translated as “those who have gone.” Although archaeologists have postulated many reasons for the Hohokam demise—climate change, invasion, flooding, environmental degradation,
Figure 7.36 An example of red-onbuff Hohokam pottery. This small jar (14 cm high) has red spirals painted on a buff-colored background, a common design element.
_ Site Snaketown
303
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 304
Figure 7.37 Hohokam artisans made a wide range of ornaments from seashells, including this shell with an etching of a horned toad (diameter 10.2 cm).
political collapse—no single factor seems sufficient to explain this abandonment. Yet the Hohokam left a lasting imprint on the desert landscape. Their ability to master the desert is testimony to the sophistication of Hohokam engineering skills. This fact was not lost on the early European settlers, who reexcavated and used some of the long-abandoned Hohokam ca-
Figure 7.38 Hohokam stone palette in human form that was used to grind pigments.
nals. Several canals still in use today in Phoenix follow the route and grade of the ancient waterways.
Figure 7.39 The network of Hohokam irrigation canals along the lower Salt River in the Phoenix Basin of central Arizona. Today Phoenix covers the area where these canals were situated; only a few of the ancient canals are still visible on the surface.
ve Salt Ri
r
Arizona Salt R. Gila R.
_ _
Snaketown
304
Chapter 7 Native North Americans
N
Large Hohokam village 0
5 km
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 305
Concept
<
Studying Community Plan at Snaketown Household clusters at Snaketown
Over the past three decades, archaeologists have paid increasing attention to the analysis of village or site plans. These studies, which require the excavation of broad horizontal areas to expose sets of associated dwellings and features, are a means of examining settlement patterns at the scale of the community. The size and spacing of houses, as well as other archaeological features—storage facilities, cemetery areas, and ritual structures, for instance— can provide researchers with information on ancient household size, the nature and arrangement of social units above the level of the household, and the association of households with storage and other social and economic activities. In the American Southwest, a long history of archaeological community plan analysis extends back to the broad horizontal excavations conducted at pithouse villages along the Mogollon Rim by Paul Martin, then at the Field Museum. In a classic study, the late James Hill used architectural and artifactual indicators to identify differences in room function at Broken K Pueblo, a late prehistoric Mogollon settlement in eastern Arizona. More recently, David Wilcox and his colleagues have made a detailed study of ancient community structure at Snaketown. Wilcox began by identifying houses that were occupied contemporaneously. He then mapped the location and orientation of each set of houses that were inhabited during the same phase. Wilcox noted when the doorways of contemporaneous structures opened onto a shared courtyard (Figure 7.40). He found that house clusters from the early occupations at Snaketown usually consisted of only two associated structures. Later residential clusters at the site had as many as six houses around a plaza. Based on
his analysis, Wilcox suggests that the people residing in each cluster of houses formed an extended family group and possibly shared certain domestic activities. Most of the Snaketown houses have similar internal features, such as small hearths near the entrance. Other domestic activities were conducted in the shared outdoor space. Outside work areas included several types of hearths, as well as postholes that seem to mark the former position of windbreaks or brush kitchens. The inside hearths were too small to hold pieces of wood large enough to produce adequate coals for cooking, which may have been done outside on a cooperative basis. Houses within each cluster were arranged at right angles to one another, defining rectangular courtyard areas. In general, clusters that contained the most houses also had the biggest courtyard areas. Over time, the location and orientation of the houses within a courtyard area showed considerable continuity, suggesting that residential placement was not random. Most larger household clusters included one structure that was square, as opposed to the usual oblong or rectangular shape. The square structures may have been used for special or ritual activities. After A.D. 900, the largest clusters of houses at Snaketown were situated at the core of the site near major platforms and a central plaza. These larger house clusters contained more total floor area than the house clusters at the edge of the site. The emergence of these large clusters, possibly signifying increased differentiation between family groups, occurred in conjunction with a greater emphasis on civicceremonial construction. Some scholars suggest that these are indicators for increasing sociopolitical complexity.
One objective of community studies is the examination of cohort groups at the residential level. Such units are important components (building blocks) within the social, settlement, and subsistence systems, and are for that reason a significant focus of archaeological research. —Glen Rice (1987)
N
Possible courtyards Houses
0
10 20 m
Figure 7.40 The distribution of pithouses and courtyards in one area of Snaketown, around A.D. 900–1100. Houses sharing courtyards were occupied by co-residential groups or families who cooperated in certain domestic activities.
Concept Studying Community Plan at Snaketown
305
_
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 306
Site
:
Chaco Canyon A prehistoric regional center in the American Southwest little permanent water. The bleak environment lacks trees and experiences dramatic temperature extremes. Rainfall, characterized by infrequent summer cloudbursts that cause washes to fill temporarily with water, is marginal for farming. Today the area is seldom used for agriculture, but once it was a center of prehistoric settlement and long-distance exchange. The Native Americans depended on rainfall for floodwater farming, in which they channeled seasonal runoff to agricultural fields, to supplement the water supply.
Peñasco Blanco Casa Chiquita Kin Kletso
Utah
New Alto Pueblo Alto
Ch
sh Wa
oW Chac
aco
Figure 7.41 A map of the Colorado Plateau in northwestern New Mexico, showing the location of Chaco Canyon and the principal Chaco roads. The early village of Shabik’eshchee preceded the construction of apartmentlike “great houses” on the floor of the canyon.
h
For a Web-based activity on the disappearance of Ancestral Pueblo culture, see the Internet exercises on your online learning center.
The Four Corners region of the Southwest—the junction of Arizona, New Mexico, Colorado, and Utah—is dominated by an extensive highland (elevation above 1500 m, 5000 ft) called the Colorado Plateau. The plateau is drained by the Colorado River and its tributaries, which have cut a complex topography of mesas, buttes, valleys, and canyons into the landscape. Located in a remote part of northwestern New Mexico, Chaco Canyon is one of the largest of these erosion features (Figure 7.41). Chaco (CHAH-ko) Canyon, 15 km (9 mi) long, has a sandy bottom and
as
w w w. m h h e . c o m / p r i c e i p 6 e
Pueblo Bonito Chetro Ketl
Hungo Pavi
Pueblo del Arroyo Casa Rinconada
Colorado
Una Vida
New Mexico
Tsin Kletzin
Wijiji
Chaco
Colorado Plateau
0
5 km
Shabik’eshchee
Chaco Canyon
Arizona
N
Chacoan communities outside the canyon Chaco roads documented by ground surveys Chaco road segments known only from aerial photographs
_ _
306
Chapter 7 Native North Americans
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Several of the most spectacular ruins anywhere in the Southwest are located in Chaco Canyon. Although brief reports of these ruins extend back to the mid-seventeenth century when military forces from Spanish outposts entered the area, the first substantial accounts were not published until 1850. At that time, James Simpson, an officer in the Army Topographical Engineers, was sent to the canyon area to investigate claims that the Navajo were harassing isolated farms and ranchos (settlements consisting of only a few houses). Simpson was overwhelmed by the massive walls, which still stood three to four stories high and formed complexes of contiguous rooms (Figure 7.42). He described and measured many of the prehistoric structures, and many other curious visitors soon followed. In 1906, Chaco Canyon was made a national monument to protect the ruins from destruction by treasure hunters and vandals. The continuity of past and present Native American groups in the Southwest is most evident between the prehistoric Ancestral Pueblo and the contemporary Pueblo Indians. Many Pueblo peoples consider the Colorado
Page 307
Plateau to be their traditional homeland. Ancestral Pueblo is now the term used for one of the major prehistoric southwestern cultural traditions that developed out of the hunting-andgathering adaptation around 2000 years ago. Before that time, temporary campsites and kill sites were present on the Chaco Plateau, but not in the canyon itself. The first sedentary Ancestral Pueblo sites date to around A.D. 100, located on mesas away from the bottomlands of the canyon, possibly for reasons of defense. Most of these sites contained 5–10 shallow pithouses. Over the next several centuries, the diet at most Ancestral Pueblo sites showed an increasing reliance on domesticated corn, beans, and squash. Yet the collecting of pine, or piñon, nuts and the hunting of cottontail and jackrabbits, deer, antelope, and Rocky Mountain bighorn sheep remained important. By the middle of the first millennium A.D., villages, which may have been inhabited year-round, were located on the floor of Chaco Canyon. Once the canyon was occupied, its population grew, and though small sites remained common, some larger communities, with 50–100 pithouses, were established.
Figure 7.42 Chetro Ketl, one of the great houses built on the floor of Chaco Canyon. For the most part, the large pueblo of Chetro Ketl was constructed during the late tenth and early eleventh centuries A.D. The outer rooms, which faced the courtyard, were used primarily for residential activities; the inner rooms were used for storage.
floodwater farming A method of farming that recovers floodwater and diverts it to selected fields to supplement the water supply.
Site Chaco Canyon
307
_
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Figure 7.43 A small kiva with a central fire pit and masonry walls.
Figure 7.44 A black-on-white Ancestral Pueblo pottery vessel. This pitcher, 17.4 cm high, is decorated with characteristic geometric designs and hatching.
kiva A semisubterranean ceremonial
room found at sites throughout the American Southwest. pueblo A stone-masonry complex of
adjoining rooms found in the American Southwest.
_ _
dendrochronology The study of the annual growth rings of trees as a dating technique to build chronologies.
308
Page 308
Kivas, semisubterranean ceremonial rooms, were constructed at some sites (Figure 7.43), and the distinctive blackon-white Ancestral Pueblo pottery first appeared (Figure 7.44). Shabik’eshchee (shuh-BIK-eh-she) is an early village in the Chaco region. Occupied between A.D. 550 and 750, the settlement had as many as 68 pit dwellings, a large kiva, numerous outdoor storage pits, and two large refuse heaps. Nevertheless, it is not clear that all these features were occupied or used simultaneously. At Shabik’eshchee, the sides of the pithouses were lined with stone slabs and coated with mud plaster. The roofs were supported by four posts topped with crossbeams. The upper walls of the houses were formed by leaning poles and sticks against the crossbeams to the ground surface outside the excavated pit. The walls were coated with a mixture of mud, twigs, and bark. One entered the house through a small antechamber. Inside the main chamber of the house was a centrally located hearth. The kiva was larger than the pithouses, and unlike the pithouses, the interior kiva walls were encircled by a low bench. Sometime after A.D. 700, aboveground rectangular rooms of adobe or roughly layered masonry were constructed on the Colorado Plateau. At first these structures may have been used for storage, while pithouses continued to serve as dwellings. In later communities, pithouses were used largely as kivas, while most dwellings were of adobe bricks and placed aboveground. The Ancestral Pueblo were the first in the Southwest to build compact villages of contiguous rectangular rooms, with different areas for habitation, storage, and ceremonial activities. For the most part, these later communities were situated at lower elevations, closer to the bottomlands, perhaps reflecting an increasing reliance on agricultural resources. After A.D. 850, there was a considerable shift in settlement patterns as population congregated in larger, apartmentlike pueblos, stone-masonry complexes of adjoining rooms. Clusters of
Chapter 7 Native North Americans
adjacent rooms served as the residences for separate families or lineages. Chaco Canyon supported at least nine large towns, or “great houses,” of several hundred rooms each, plus hundreds of smaller villages of 10–20 rooms each. The towns were large, multistory complexes, standing as high as four stories at the back wall, and averaging 288 rooms, some of which were large with high ceilings. The towns appear to have been built according to a preconceived plan, and there was a high degree of quality and uniformity in masonry styles. Each town also had at least one “great kiva.” Placed in interior courtyards, these large kivas had a central ceremonial role in community activities. Many smaller kivas were located elsewhere in the towns and most likely served as ceremonial chambers for the various kin or family units that made up the site population. These larger sites served as local centers for resource redistribution, long-distance trade, and ceremonial activities. The largest and most impressive town, Pueblo Bonito, covers more than 1 ha. It is a huge, D-shaped complex composed of over 600 rooms arranged in several stories (Figure 7.45). When Pueblo Bonito was at its height, as many as 500–1000 people may have resided there, although archaeologists disagree on the number of rooms that were residential in character and the proportion of the rooms that were occupied contemporaneously. Situated at the base of a 30-m (100-ft) mesa on the north side of the canyon, the pueblo is protected by the steep cliffs of the canyon. The occupation of Pueblo Bonito has been precisely dated using dendrochronology, the study of the annual growth rings of trees as a dating technique to build chronologies (Figure 7.46). This highly precise technique, pioneered by A. E. Douglass more than 60 years ago, is based on the recognition that certain trees produced distinctive ring sequences in response to shifts in temperature and precipitation. Based on the work of Douglass and his successors, a continuous treering chronology for the Colorado Plateau has been projected back more
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 309
Figure 7.45 Pueblo Bonito. The circular features are kivas. Interior rooms were used for storage, and outer rooms facing the courtyards served as living quarters.
than a thousand years. Tree-ring dates obtained from preserved wooden beams at Chaco Canyon place the earliest building at Pueblo Bonito at A.D. 861; town construction apparently was finished by A.D. 1115. At Pueblo Bonito, the masonry consisted of layers of stone covering an interior core of rock and adobe rubble. The room walls were faced with adobe plaster. The rooms surrounded central court-
yards. The outer rooms, which had doors and windows facing onto the courtyard, served as living quarters, and interior rooms were used for storage. The most elevated rooms were at the back of the complex along its outer rim. These back rooms included one great kiva and several smaller ones. This great kiva, the largest in the community, measured 20 m (65 ft) in diameter and was encircled by a wide masonry bench.
Reconstructed sequence
Figure 7.46 Dendrochronology. The matching of tree rings from a modern tree with a known cutting date to those of progressively older tree samples results in a long sequence of distinctive tree-ring patterns that can be used to date beams from archaeological sites.
Beams from archaeological sites
_
Tree stumps from a living tree
Site Chaco Canyon
309
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
The many questions that remain unanswered about the Chacoan regional system signify both the accomplishments and the challenges of Southwestern archaeologists interested in the problem of sociopolitical development. After nearly a century of survey and excavation throughout Chaco Canyon and the San Juan Basin, investigators have . . . conclude[d] that the Chacoan regional system was hierarchically organized. This conclusion has prompted three major questions . . . : How did this complex system develop? What kinds of relationships existed among its different components . . . ? Why did the system collapse? —Jill Neitzel (1989)
jet A compact, black coal that can be
highly polished; used to make beads, jewelry, and other decorative objects. macaw Any of several varieties of
_ _
parrots from Mexico and Central and South America that were prized for their colorful feathers.
310
Page 310
From A.D. 1020 to 1100, the Chacoan system peaked in population and spatial size. Much of the population of the northern Southwest participated in the regional trading network centered at Chaco Canyon. Throughout this network, which covered at least 53,000 sq km (20,500 sq mi), there were at least 125 planned towns with distinctive Chacoan architecture. Settlements were linked by a complex of roads radiating from the canyon, built in straight lines, not contoured to the topography, with ramps and stairways ascending the cliffs. Some were lined with masonry curbs, and some were up to 9 m (30 ft) wide, leading to sites up to 190 km (120 mi) away. Chacoan towns evidently had the capability of mobilizing large labor parties for construction. Timber was cut from forests up to 90 km (56 mi) away. Imported turquoise was worked in Chaco Canyon; some scholars have suggested it functioned as a medium of exchange. Such exotic goods as jet, turquoise, shell bracelets, iron pyrite mosaics, conch shell trumpets, ornamental copper bells, and macaw feathers are found more frequently at the large central town sites than at smaller villages. Social differentiation and some form of hierarchical leadership appear to have been in place, although the systems do not seem to have focused on special individuals. The nature of Chaco’s political formations remains a matter of stimulating archaeological discussion despite the many years of productive multidisciplinary research conducted in the canyon and surrounding area. The Chaco regional system was disrupted in the early 1100s. The population declined, although complete abandonment of the canyon did not occur for almost 200 years. The demographic collapse in the canyon coincided with population increases in other parts of the Colorado Plateau, including surrounding upland regions. In some areas, villages shifted to wellprotected, defensive locations, such as the sheltered cliffs at Mesa Verde in southwestern Colorado. Defense was a likely motive in the resettlement of
Chapter 7 Native North Americans
some Ancestral Pueblo villages, and this pattern may reflect an era of political instability that followed the collapse of the Chacoan centers. In roughly A.D. 1300, the Colorado Plateau region went through another major restructuring, and many Pueblo sites were again abandoned. When Europeans arrived in the Southwest, only a small number of large Ancestral Pueblo settlements remained. Whether these episodes of dispersal and reorganization (first at Chaco and later in other parts of the Ancestral Pueblo region) were triggered by climate change, environmental degradation, shifting trade connections, changing political alliances, or other factors remains a matter of conjecture and discussion. Although the causes of these abandonments are still debated, the construction and maintenance of Pueblo Bonito, a structure that was the largest apartment building erected in the United States before the nineteenth century, in a dry, desolate canyon, remains a powerful testament to the ingenuity of the Native Americans who lived there.
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 311
Site
:
Ozette A prehistoric Northwest Coast whaling village
The Pacific Coast of Oregon, Washington, and British Columbia is an environmentally rich area where land and sea hold a wealth of natural resources. The sea and the rivers contain mollusks and many species of fish, including salmon, halibut, cod, and herring. Sea mammals—such as seal, sea lion, otter, porpoise, and whale—thrive in the offshore waters. Waterfowl can be found along the shore, and farther inland there are deer, elk, bear, and smaller animals. The area has a heavy forest cover of fir, spruce, cedar, and some deciduous trees. Although vegetable foods are less plentiful, many species of berries abound. The earliest known inhabitants of the Pacific Northwest were mobile hunter-gatherers who moved into the area before 8000 B.C. At first, coastal foods did not constitute a major portion of the diet, but by 2000–3000 B.C., the accumulation of large shell middens at several sites indicates an increasing use of the readily available marine shellfish. Over the past 2000 years, subsistence patterns have continued to emphasize the exploitation of fish, shellfish, and sea mammals, supplemented by land mammals and birds. Such plant foods as berries, roots, and bulbs were not staples, but they were available during lean times of the year. In later prehistory, the peoples of the Pacific Northwest were specialized hunter-gatherers who lived in permanent villages and developed large food surpluses by exploiting a variety of fish species. Although subsistence was based on wild resources, many of these societies had specialists in hunting, fishing, curing, and toolmaking. Large, permanent settlements of several hundred people appeared by A.D. 1000, despite the absence of agriculture. The abundance of giant cedar trees provided
plentiful material for building houses and making dugout canoes. The natural wealth of the environment and the range of available foods allowed for the production of surplus goods and ornate material items, such as decorative wood carvings, including totem poles, carved boxes, canoes, and masks, as well as cedar bark baskets and textiles. Canoes made long-distance travel possible, facilitating the gathering of seasonal resources, the hunting of sea mammals, and the maintenance of farreaching social networks. Social ranking appeared on the Northwest Coast by 500 B.C. At the time of European contact, all individuals in some Pacific Coast societies were ranked into a series of relatively higher and lower statuses, according to both heredity and wealth. There were chiefs and slaves and, in between, craftspeople, hunters, and fisherfolk. Most respected of all were the whale hunters. Whaling could be undertaken only by
Figure 7.47 Ozette Village, located on the outer coast of the Olympic Peninsula. Excavations of the village were situated near the beach on the left side of this photograph.
totem pole A pole or post that has
been carved and painted with totems or figures, such as animals, that serve as the emblems of clans or families.
Site Ozette
311
_
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Figure 7.48 A carved wooden box from Ozette. The wet conditions at the site preserved objects of wood and fiber that usually do not preserve well in archaeological contexts.
_ _
312
Page 312
men from wealthy families; it was a hereditary right. Only chiefs possessed the necessary wealth to build whaling canoes, to outfit them, and to assemble the crew. Hunting whales in these large, oceangoing canoes was very dangerous. Bringing home one whale brought enormous prestige to a family, in addition to vast amounts of food. One well-known Pacific Northwest settlement, Ozette (oh-ZEHT), is located on the coast of Washington’s Olympic Peninsula at Cape Alava, the westernmost point of the contiguous 48 states (Figure 7.47). Because the cape juts out into the Pacific Ocean, Ozette is close to the migration routes of a variety of whales, including gray, humpback, and sperm whales. The shoreline is a crescent-shaped beach protected by points of land, an offshore reef, and small islands. Abundant sea resources and a rich forest behind the beach made the cape an attractive location for prehistoric Native American groups. People settled in the area over 2000 years ago, and eventually Ozette grew into a major whaling village. Ozette is unique in that it was one of a few major whaling villages south of Alaska. Two other sites, south of Ozette on the coast, are located near quiet bays. Their long, straight beaches are washed by a rolling surf, making canoe travel much more difficult than at Ozette. In addition, sea mammals pass too far offshore to be worth hunting. As a consequence, those two sites show less reliance on ocean resources. Rather, the inhabitants focused on the land and bay resources, such as deer,
Chapter 7 Native North Americans
elk, harbor seals, and salmon, in addition to oysters, mussels, and clams. The village of Ozette stretched for almost a mile along the coast and had a maximum population of roughly 800 people. It is thought to have been a major settlement of the Makah Indians, a group that still resides on the Olympic Peninsula today. Although the population of Ozette declined after Europeans began to settle in the area during the mid-nineteenth century, it continued to be occupied until the late 1920s. Today, there are few surface remnants of the village left. Richard Daugherty, formerly of Washington State University, began excavations at the site in the late 1960s when told of native accounts of mudslides that were believed to have periodically buried parts of the site. However, the actual discovery of the buried village area occurred accidentally after a violent winter storm in 1970. The harsh waves of the storm eroded sections of a bank and exposed a number of timbers, baskets, boxes, paddles, and other wooden artifacts, which were discovered by hikers. That spring, Daugherty began excavation of the buried houses. Part of Ozette has been preserved by a massive mudslide that buried five houses around A.D. 1500. Because the heavy layer of clay sealed the houses and kept out oxygen, perishable artifacts of wood and fiber, which normally do not preserve well in the wet environment of the Northwest Coast, could be recovered (Figure 7.48). Since the village was occupied at the time of the mudslide, the excavated material
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 313
Figure 7.49 Examples of fishhooks made from bone and wrapped with bark. The abundance of fishhooks and points recovered during excavations at Ozette is an indication of the importance of marine resources to the site’s inhabitants.
represents the entire range of wooden artifacts that were used by a Northwest Coast household. In addition to the wooden planks and posts of the houses, the excavators recovered baskets, mats, hats, tumplines, halibut hook shanks, arrow shafts, harpoons, finely carved wooden clubs and combs, box fragments, bowls, wood wedges, and a variety of fishhooks and barbed fish points (used on the end of a spear to stab fish in the water) (Figure 7.49). To date, over 60,000 artifacts have been recovered. As a result of the region’s abundant rainfall and high groundwater, Ozette is water-saturated, like many other sites on the Northwest Coast. This condition has made normal excavation procedures difficult. When trowels and shovels were used, as they are on most excavations, they sliced through and gouged the fragile wood and fiber. To deal with this problem, a method called wet-site excavation was developed. With this technique, water is pumped through garden hoses and sprayed onto the deposits to remove the dirt and expose the archaeological materials. Excavators use high water pressure to remove the heavy clay deposits and low pressure to remove dirt from more fragile artifacts. This procedure works well because the water pressure can be adjusted continually to expose an artifact without dis-
locating or destroying it. By using a very fine spray, excavators can carefully reveal and remove even the remains of basketry and other fibers (Figure 7.50). The excavated houses at Ozette were very large, some 20 m (60 ft) long and 10 m (33 ft) wide, about the size of a tennis court. They were constructed of cedar planks, some up to 0.5 m (1.5 ft) wide, supported by upright wooden posts, and held in place with twisted cedar twigs (Figure 7.51). Roof boards were overlapped to keep out the rain. Raised platforms ringed the inside walls and were used for sleeping and storage. Most of the recovered artifacts were found in association with these platforms. The existence of several hearths in each house suggests that these large structures were occupied by more than one nuclear family. In one house, the highest-ranking family apparently lived in the left-rear quarter; ceremonial gear and whaling harpoons were found there. In this area, a woven cedar bark hat, traditionally worn by individuals of high status in later times, also was recovered. In another house, wood chips associated with woodworking activities were found. Excavations at Ozette have provided ample evidence for the extensive hunting of sea mammals. Whale bones were found in the earliest levels, indicating that the prestige and
Figure 7.50 A basket woven of cedar bark. Cedar bark was used to make a range of baskets for collecting and storing food, textiles including blankets and conical rain hats, and sleeping mats.
tumpline A strap that is passed over the forehead or the chest to facilitate the transportation of a heavy load carried on the back. wet-site excavation The technique
of excavating waterlogged sites by pumping water through garden hoses to spray the dirt away and expose archaeological features and artifacts.
Site Ozette
313
_
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 314
Figure 7.51 An artist’s interpretation of a house at Ozette. The walls and roofs were constructed of split cedar planks.
Figure 7.52 Petroglyphs of whales and human faces at Wedding Rock near Ozette.
_ _
petroglyph A drawing that has been carved into rock.
314
practice of whaling date back at least 2000 years. Many whale bones also were found in later levels; some of the bone was used in house foundations, to shore up walls, or to divert water and mud from sliding down the hillside onto the backs of the structures. At low tide, a cleared strip through the offshore rocks is still visible. Centuries ago, canoes could have taken this route to the open waters to encounter the migrating whales. Artifacts associated
Chapter 7 Native North Americans
with whaling—parts of canoes, canoe paddles, and whaling harpoons—were plentiful at the site. The people of Ozette also hunted other sea mammals, including fur seal, porpoise, dolphin, and sea lion, which constituted the majority of mammal bones at the site. Elk and deer bones were present, but in small numbers. Although salmon bones were few at Ozette, halibut, ling cod, and shellfish were well represented. The importance of maritime resources is also evident 3 km (2 mi) south on the beach at Wedding Rock, where at least 44 petroglyphs have been found on beach boulders. Whale motifs are especially prevalent, including one of a pregnant whale (Figure 7.52). Although a rich ceremonialism on the Northwest Coast was associated with the hunting of sea mammals, especially whales, there is no definitive evidence that the petroglyphs were used as part of such ceremonies. One possible function of the carvings may have been to gain supernatural power over the inhabitants of the sea. Whale hunters may have made the petroglyphs to ensure good luck on whaling trips or to commemorate other important events. No method of dating the petroglyphs is yet available. However, the Wedding Rock art may have been made at the same time that the excavated houses at Ozette were inhabited, because several of the carved motifs are also present on artifacts recovered from the site.
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 315
Concept
<
Chiefs Hereditary leadership among native North Americans
The first Europeans on the North American continent encountered a wide variety of social and political organizations among Native Americans. Some groups were organized as bands or tribes, with little formal variation in the status or rank of individuals. People filled different roles according to their talents and achievements. At the other end of the spectrum were a number of groups in which positions of authority were inherited at birth. At the top of this hierarchy of social rank were chiefs, who held political authority over individuals in the society. Native American chiefs engaged in a wide range of activities, including intersocietal politics, ceremonial leadership, and declarations of war (Figure 7.53). Yet none of these chiefly societies was as hierarchically organized or socially stratified as the largest polities to the south.
In the Americas before European contact, chiefs usually were associated with agricultural societies. Such was the case for chiefs present in many parts of Central America and the circumCaribbean, as well as in some areas of North America. Although food surpluses can be produced by most, if not all, economies, it often is easier to produce a storable surplus with agriculture. Domesticated plants are generally easier to store than wild resources, and they tend to produce a greater return per unit of land. Chiefs can extract or control food surpluses; they also can encourage their kinsmen to work to produce beyond their immediate needs. In the southeastern United States, such large sites as Cahokia and Moundville were located in very fertile areas. The mortuary mounds and other burial contexts where select individuals were accompanied by elaborate
The evolution of complex societies is one of the most interesting questions in all the social sciences. How can a species long adapted to living in small egalitarian groups evolve revolutionary new social institutions that lead them to live in very large, highly inegalitarian social systems? —Peter Richerson and Robert Boyd (2001)
Figure 7.53 A sixteenth-century engraving of Mississippian chief Holata Outina surrounded by his troops as he leads them into battle against his enemies.
_ Concept Chiefs
315
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 316
Figure 7.54 A carved, wooden potlatch mask from the nineteenthcentury Northwest Coast.
y
The emergence of institutionalized social inequality was a fundamental human transition that apparently occurred relatively independently in many different places. Can you think of some reasons why? How did this important organizational change occur? How can different ideas about this transition be evaluated?
potlatch A large feast among North-
west Coast Native Americans that included the display and dispersal of accumulated wealth to the assembled guests.
grave offerings indicate the presence of powerful chiefs at these sites. Chiefs also were present in some societies that did not practice agriculture. Most societies on the Northwest Coast were based on hunting and gathering. Their economies focused on a rich and diverse base of marine resources, which permitted the accumulation of surplus. These societies also had hierarchies of social rank. Only people from the highest-ranking families could become leaders. Chiefs did not necessarily return all the produce they collected. This surplus often provided the resources to obtain exotic goods through trade. Much chiefly exchange involved the acquisition of high-status items that were either traded to establish new al-
_ _
316
Chapter 7 Native North Americans
lies or loaned to attract followers. Often a chief’s power was measured by the number of allies and dependents that could be mobilized for specific tasks. Although chiefly positions were generally acquired through familial ties, personal achievements also were important. Among Northwest Coast Native American groups, only highranking individuals could attain the position of chief. But one’s role as chief had to be validated continually; failure to do so resulted in a loss of status. One way chiefs validated their position was by publicly displaying their wealth and distributing their accumulated property. At the time of European contact in the Northwest, an elaborate ceremonialism had developed, based on the potlatch, a large feast that included the display and dispersal of accumulated wealth to assembled guests (Figure 7.54). The mere possession of wealth did not confer prestige. But by distributing the wealth, the chief created social debts and obligations that could be called in and used to bolster his position at a later date. A potlatch gave the host far greater prestige than selling or trading, but in economic terms, the result was relatively similar: the redistribution of goods. Politically, the quest for prestige gave momentum to the whole system; the more a chief gave away, the greater his status. To date, most archaeological interpretations have emphasized the functional or managerial advantages, especially the mitigation of food crises, associated with the emergence of chiefs. Yet such arguments are incomplete; in many cases, the likelihood of major crop failures requiring chiefly intervention was low. The construction of complete explanations requires that archaeologists extend consideration beyond the benefits of chiefly societies to the political strategies of emergent leaders and the factors that allowed those strategies to work for the benefit of few at the expense of many.
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
Page 317
Images and Ideas The Clash of Worlds Changing perspectives on native North Americans
W
hen European settlers and slaves from Africa first penetrated the Americas north of Mexico in the sixteenth, seventeenth, and eighteenth centuries, they observed huge earthen mounds at sites in the East that were no longer inhabited, as well as well-planned mud-brick ruins in the West. Blinded by the ethnocentrism of the era, many of the early explorers failed to recognize any of the obvious historical connections between the land’s indigenous peoples and the impressive architectural features. Frequently, the Euro-American traders and adventurers speculated that the great earthworks and pueblos were remnants of earlier constructions built by Romans, Vikings, Celts, or people from imaginary lands, such as Atlantis. Few early explorers were willing to accept the possibility that the ruins were part of the heritage of Native American peoples, whose lands and resources they coveted. Sadly, such migrationist views are still all-toofrequently advanced (and may even attract occasional popular attention today), although they remain entirely without solid empirical support. (Some authors have gone so far as to postulate extraterrestrial contacts!) Fortunately, there were exceptions, such as Thomas Jefferson. In the nineteenth century, a principal focus of early North American anthropology was the Native American, in the past as well as in the present. Information was collected from archaeological evidence, early historical accounts, and contemporary observations. Continuities in language, ritual, and material culture were emphasized in an effort to confirm the historical relationships between living indigenous peoples, past documentary records that described the aboriginal inhabitants of North America, and artifact inventories. Although many of the associations between present and past were reasoned and justified (laying the foundation for twentieth-century archaeology and anthropology), the absence of an adequate time scale (no absolute dating techniques were yet invented) contributed to the occasional overreliance on contemporary or historical records to interpret the archaeological past. Archaeology depends on inspirations, clues, analogies, and models from more recent times to help flesh out the past. Yet archaeologists studying the prehistory of North America, as well as other areas, should be prepared to recognize changes that occurred during prehistory and immediately thereafter. Prehistoric peoples may have had a way of life and organizational formations that were markedly different from those recorded in written texts. Many studies have documented that the contact period, the sixteenth through eighteenth centuries, was a time of great change for most, if not all, native North American groups. In many areas, such as the Southeast, Native Americans not only were savaged in combat with the European invaders (such as the nine shiploads of conquistadores, led by Hernando de Soto, who landed on the coast of Florida in 1539) but also were vanquished by infectious, epidemic diseases (influenza, smallpox, measles, and whooping cough) introduced to the Americas. Because the aboriginal peoples had no immunity to these diseases, their effect often was calamitous, decimating the native populations. In many regions, the impact of disease may have been further intensified by social and economic dislocations occurring at the same time. Regardless of the specific causes, many Native American populations had been severely ravaged and disrupted in demographic size, subsistence, and sociopolitical organization when they first were encountered by more permanent settlers (and eventually anthropologists) during the seventeenth, eighteenth, and nineteenth centuries. As a consequence, anthropologists and
< ( : < V : y < p : W
ethnocentrism Evaluating other groups or societies by standards that are relevant to the observer’s culture.
Images and Ideas The Clash of Worlds
317
_
pri31057_ch07.pp266-319.qxd
10/1/09
8:07 AM
The history of European colonial expansion following the late 15th century is riddled with a multitude of curious and seemingly inexplicable encounters between native cultures and Europeans, which demand the conjoining of historical and anthropological methods. —Jeffrey L. Hantman (1990)
The disadvantage of men not knowing the past is that they do not know the present. History is a hill or a high point of vantage, from which alone men see the town in which they live or the age in which they are living. —Gilbert K. Chesterton (1933)
_ _
318
Page 318
archaeologists must be careful about relying too specifically on direct analogies between ethnohistoric, as well as early ethnographic, accounts and the deeper past that we see through archaeology. Nevertheless, many aspects of tradition, belief, and religion do show significant continuities between precontact and postcontact times. When the Europeans first arrived in North America, few, if any, sites as large as Cahokia in the Midwest or the pueblos of Chaco Canyon in the Southwest appear to have been inhabited, or were described. Yet this fact should not lead us to doubt the Native American heritage of these sites or force us to inhibit our interpretation of the past by assuming that the archaeological settlements were organized in exactly the same way as those of the ethnographically observed inhabitants of the respective regions. Such variation should not be surprising, since diversity in both space and time seems to have been a key feature of the Native American way of life. In part, spatial variation may have been related to the great environmental differences that characterize the North American continent. At times, cultural changes may have been responses to documented episodes of climatic transitions. Yet such environmental factors alone cannot explain the range of prehistoric variability. Perhaps the smaller and less stratified social systems that occupied North America were generally more flexible and organizationally (and demographically) fluid than the more complex, hierarchical polities that developed in much of Middle and South America (see Chapters 8 and 9), as well as in a good portion of the Eastern Hemisphere (see Chapter 10). Such flexibility and fluidity also may help account for the great cultural diversity of the prehistoric North American peoples. Although there still are gaps in our knowledge, anthropological archaeology has established that Native Americans inhabited vast tracts of North America for at least 12,000 years. It is clear that they adjusted to and modified the highly diverse environments of North America through a broad range of technological, social, and cultural behaviors and strategies. Archaeologists have documented that native lifeways changed over time, often gradually, but sometimes much more rapidly. In this chapter, we have reviewed key settlements, cultures, and lifeways of the native peoples who inhabited North America before the influx of peoples from the Eastern Hemisphere, or “Old World,” during the latter half of the second millennium A.D. This chapter’s discussions have drawn on more than a century of archaeological study, in conjunction with investigations in related fields. This research has illustrated how native North Americans thrived in locales as diverse as the deserts of southern Arizona, the woodlands of eastern North America, and the coast of the Pacific Northwest. Our review also has highlighted how the prehistoric peoples of North America established a diverse array of hierarchical social formations. Some of these had clear hereditary chiefs, but others appear to have had forms of social differentiation that focused less on individual leaders or special personages (such as at Chaco Canyon during much of its history). Interestingly, although many kinds of hierarchical leadership developed in North America, the kinds of writing systems, institutionalized bureaucracies, and large urban centers that arose in many other regions had not materialized in this cultural region before 1492. Nevertheless, similar to what archaeological findings have shown for humans around the globe, the behaviors and lifeways of native North Americans were far from entirely stable or unchanging. Both enduring traditions and significant transitions were key aspects of their long-term history even before the tragic disruptions that followed European contact. For this reason, it is important to recognize that archaeological data have a significant role to play in liberating and amplifying Native American history from an exclusive reliance on documentary sources, which are primarily products of Euro-American culture and pertain directly only to the end of the precontact era.
Chapter 7 Native North Americans
pri31057_ch07.pp266-319.qxd
p
10/1/09
8:07 AM
Page 319
w w w. m h h e . c o m / p r i c e i p 6 e
DISCUSSION QUESTIONS 1. What are the major differences between reciprocity and redistribution? How might you distinguish these exchange practices archaeologically?
For more review material and study questions, see the self-quizzes on your online learning center.
2. What other systems of exchange can you think of? 3. How can burial behavior be informative about the nature of social/economic organization in the past? 4. What are the major similarities or differences in the nature of settlement layout between the eastern Woodlands and the North American Southwest? 5. What do you think these similarities or differences can tell us about past social, political, and/or economic organization? 6. What is the nature of the relationship between chiefly organization, surplus production, and subsistence? 7. Why did Eurasian diseases have such a calamitous effect on indigenous Americans?
z
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e.
Cordell, L. S. 1997. Archaeology of the Southwest, 2d ed. San Diego: Academic Press. The best current synthesis of archaeology in the American Southwest. Crown, P. L., and W. J. Judge, eds. 1991. Chaco and Hohokam: Prehistoric regional systems in the American Southwest. Santa Fe, NM: School of American Research. A scholarly compendium of papers that compare and contrast regional variation in southwestern prehistory. Fagan, B. 2005. Ancient North America, 4th ed. London: Thames & Hudson. An amply illustrated synthesis of North American archaeology. Feder, K. L. 2001. Frauds, myths, and mysteries: Science and pseudoscience in archaeology, 4th ed. New York: McGraw-
Hill. An entertaining and informative exploration of fascinating frauds and genuine mysteries that relate to the human past. Gibbon, G., ed. 1998. Archaeology of prehistoric Native America: An encyclopedia. New York: Garland. A comprehensive compendium of short essays concerning key topics and sites in North American archaeology. Neitzel, J. E., ed. 1999. Great towns and regional polities in the prehistoric American Southwest and Southeast. Albuquerque: University of New Mexico Press. A series of scholarly essays that compare pre–European contact social formations in the North American Southwest and Southeast.
_ Suggested Readings
319
pri31057_ch08.pp320-389.qxd
C
10/1/09
H
8:11 AM
A
Page 320
P
T
8 E
R
E
I
G
H
T
Figure 8.1 Reconstruction of the Aztec Templo Mayor superimposed on the modern buildings in Mexico City that were constructed on top of the ruins of the razed temple.
_ _
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 321
Ancient Mesoamerica Introduction Early State Development in Mesoamerica The Aztecs, the Maya, and their neighbors
I
n the same year, 1492, that Christopher Columbus landed in the Caribbean, a boy is believed to have been born in a Spanish village, Medina del Campo, Castile. Decades later, that boy, Bernal Díaz, then a young man, traveled to Mexico in the company of Hernán Cortés, and later in life he recorded his memories. Díaz’s written texts provide a gripping firsthand account of the clash of two great powers, sixteenth-century Spain and the Aztecs of central Mexico, a confrontation that led to the construction of the churches and government buildings of the conquering Spanish immediately on top of the ruins of the earlier Aztec temples and palaces (Figure 8.1). Díaz’s recollections of entering the Aztec island capital, Tenochtitlán, are particularly poignant: (W)hen we saw so many cities and villages built in the water and other great towns on dry land and that straight and level causeway going towards Mexico, we were amazed and said that it was like the enchantments they tell of in (legend), on account of the great towers and cues and buildings rising from the water, and all built of masonry. (1956:190)
Later, Bernal Díaz wrote of the great market in Tenochtitlán’s sister city, Tlatelolco: (W)e turned to look at the great market place and the crowds of people that were in it, some buying and others selling, so that the murmur and hum of their voices and words that they used could be heard more than a league [three miles] off. Some of the soldiers among us who had been in many parts of the world, in Constantinople, and all over Italy, and in Rome, said that so large a market place and so full of people, and so well regulated and arranged, they had never held before. (1956:218–219)
These descriptions make it clear that the settlement patterns and economic organization of the late prehispanic Aztecs were rather different from any of the social formations that emerged north of Mexico before the arrival of Europeans. The Aztecs were a state, and through conquests of neighboring peoples and polities (some of whom were ethnically and linguistically very different), they had begun to establish an empire. Archaeologists and other social scientists define a state as an internally specialized and hierarchically organized political formation that administers large
< ( : < V : y < p : W
CHAPTER OUTLINE
Introduction Early State Development in Mesoamerica 321 SITE San José Mogote
326
CONCEPT Nonresidential Architecture 329
SITE San Lorenzo and La Venta CONCEPT The Olmec Horizon SITE El Mirador
337
CONCEPT Carved Stones and Early Writing 341 SITE Monte Albán
342
331
335
CONCEPT Settlement Pattern
Surveys 347 SITE Teotihuacan
349
CONCEPT The Mesoamerican Ballgame 356 SITE Tikal
357
CONCEPT Tikal’s Monument
Record 361 CONCEPT Wetland Fields SITE Palenque
365
363
CONCEPT Writing and Calendars SITE Tula
371
SITE Chichén Itzá
374
SITE Tenochtitlán
378
CONCEPT Aztec Markets
383
CONCEPT Human Sacrifice
and Cannibalism 384
Images and Ideas The End 321
369
of Prehispanic Civilizations in Mexico 386
_
pri31057_ch08.pp320-389.qxd
10/1/09
w w w. m h h e . c o m / p r i c e i p 6 e
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
state A form of government with an internally specialized and hierarchically organized decision-making apparatus. A state generally has three or more administrative levels. empire A union of dispersed territories, colonies, states, and unrelated peoples under one sovereign rule. estuary A low area along a coast
where the wide mouth of a river meets the sea and the waters of the two mix.
8:11 AM
and complex polities. States are associated with populations that are socially and economically stratified. In these societies, wealth, social status, and political power generally are inherited, whereas in unstratified societies, such advantages are most often achieved through personal skills and experience. Although some archaeologists have suggested that Cahokia, in the midwestern United States, was at the center of a short-lived state organization (see Chapter 7), most researchers argue that state formation in the Western Hemisphere did not occur north of Mexico before European contact. Indigenous or pristine states—institutions that developed relatively free of significant external contact or outside influence—arose in Mesoamerica, South America, Southwest Asia, and China. The rise of these early polities, and several cases of secondary and tertiary state development, are presented in this chapter, as well as in Chapters 9 and 10. In this chapter (Mesoamerica) and Chapter 9 (South America), we take a more historical approach in surveying the complex processes surrounding the rise of civilizations in two areas of the Western Hemisphere. Although comparisons are made throughout, we adopt a more deliberately comparative perspective in Chapter 10, where we discuss early civilizations from several continents in the Eastern Hemisphere. Chapter 10 concludes with a discussion of some of the theories that have been advanced to explain state origins. The first indigenous states were established in Southwest Asia during the fourth millennium B.C. For that reason, some of you may wish to review Chapter 10 before reading Chapters 8 and 9. Because so many of the theories about state origins were constructed with Southwest Asian cases in mind, we think it is helpful to consider these different explanations in Chapter 10 in conjunction with the examples from the Eastern Hemisphere. We also want to discuss these theoretical positions after we have considered the variability of early states. For that reason, we begin our discussion with Mesoamerica, drawing comparisons between the highland polities (such as Teotihuacan) and those of the Maya Lowlands. We then move to Andean South America in Chapter 9, where we consider some of the similarities and differences between the prehispanic Andean and Mesoamerican worlds. In Chapter 10, we offer a more global, synthetic perspective. When Cortés and his Spanish conquistadores landed on the eastern coast of Mexico in A.D. 1519, they encountered a remarkably diverse landscape of cultures and environments. The terrain is a complex vertical mosaic of snow-capped volcanic peaks and arid highland valleys, lush tropical forests, “scrubby” plains, swampy lowlands and estuaries, and beautiful sand beaches. The climate ranges from the arctic cold of high mountain summits to the sweltering heat at sea level, and climatic and topographic variability results in a rich mosaic of animal and plant life. Before the 1519 arrival of the Spanish, Mesoamerica—which includes central and southern Mexico, the Yucatán peninsula, and the northern parts of Central America—was also a world of enormous linguistic and ethnic differences. The prehispanic inhabitants spoke many languages, employed a wide range of farming and water-control strategies, and inhabited towns and cities of diverse form and function. Yet the peoples of Mesoamerica also shared a great deal, including a reliance on similar staple foods, widespread trade, and related religious systems. This shared ceremonial realm included a calendar, stepped pyramids, ritual sacrifice of blood, writing systems, and specific styles of dress. The dietary “trinity” of corn, beans, and squash (supplemented by avocados and chili peppers) provided a remarkably nutritious diet. Maize (corn) provides most of the essential amino acids for building protein; lysine, a missing ingredient, is present in beans. An enzyme in squash contributes to the digestion of the protein in beans. Maize extracts nitrogen from the earth in which it grows; beans release nitrogen back to the soil. Corn and squash are rich in carbohydrates and calories. Beans and the
_ _
322
Page 322
Chapter 8 Ancient Mesoamerica
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 323
avocado provide fat. The chili pepper, a regular condiment in Mesoamerican meals, is high in vitamins and aids in the digestion of high-cellulose foods. As discussed in Chapter 6, corn, beans, and squash were domesticated by 6000–10,000 years ago in the highlands of Mexico, where the wild ancestors of these plants still grow side by side. These domesticated plants spread throughout Mesoamerica by 2000 B.C. At about that same time, perhaps a few centuries earlier, people began to make simple pottery in the shape of squashes and gourds. Yet it was not for centuries (roughly 1800–1500 B.C.) that we find the first permanent farming villages in this region. This pattern differs from what we saw in parts of North America (see Chapter 7), where some residential stability was present before reliance on domesticated plants. The sequence of development in highland Mesoamerica appears to be the domestication of plants, followed by the invention of ceramics, and then the emergence of village society. Once the indigenous peoples of Mesoamerica had adopted a sedentary farming lifeway, the advent of the region’s earliest cities and states followed after roughly 1000 years. The Spanish conquest 2000 years later effectively terminated the indigenous rule of native Mesoamerican civilizations. Information about those civilizations and their predecessors comes from a number of sources, including Spanish accounts, prehispanic texts, archaeology, and ethnography. The Spanish conquerors and priests kept diaries and descriptions of their impressions of this new world, which often incorporated the remembrances of indigenous confidants as well. These ethnohistoric sources provide an invaluable record of ancient Mesoamerican customs, beliefs, and individual histories at the time of European contact. However, they also provide an incomplete and often ethnocentric picture of the prehispanic past. Scholars also derive information from the texts of the prehispanic peoples themselves. These records include both short accounts, written more than a millennium ago on stones, murals, and pottery, and a few later (and generally longer) books, or codices. The surviving codices date principally to the end of the prehispanic era. The religious zeal of the Spanish resulted in the destruction of much that was native, including the burning of many books, the razing of temples and palaces, and the destruction of the priesthood. Decipherment of the surviving written accounts, particularly those of the ancient Maya, has greatly enhanced our understanding of rituals, militarism, and rule in prehispanic Mesoamerican civilizations. The native texts are limited in scope and geographic coverage, however, and they reflect a bias toward the events in the lives of the literate elite and rulers. Ethnographic studies of contemporary peoples in modern Mesoamerica provide information about the organization of agricultural systems and the survival of certain prehispanic customs. Nevertheless, the Spanish presence in Latin America effectively eliminated much of what was native through law, education, and religious doctrine. Despite these historical and modern sources of information, archaeological research is essential for a basic understanding of the emergence of Mesoamerican cities and civilizations. More than a century of investigations in Mesoamerica has yielded significant findings. Major tomb and temple excavations have been conducted in lowland regions of Mesoamerica, where ancient Maya ruins sat virtually undisturbed from the time of their abandonment around A.D. 900 until their rediscovery in the nineteenth century. More recent studies have analyzed villages and their component houses (household archaeology) and the hinterlands of ancient cities, rather than exclusively temples and tombs, to understand the organization and operation of everyday life. In Mexico’s Central and Southern Highlands, where archaeological preservation is generally good and ancient ruins are often visible on the present land surface, long-term, multistage archaeological excavation and settlement pattern survey programs have been implemented.
ethnography The study of human cultures through firsthand observation. codex (plural codices) (Latin) A hand-
painted book on bark paper or animal skins folded like a screen. household archaeology The archaeological analysis of past houses and associated residential remains to learn about domestic life and activities.
_ Introduction Early State Development in Mesoamerica
323
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 324
Gulf of Mexico
Chichén Itzá YUCATÁN
El Tajín
Tula
Tenochtitlán
Teotihuacan
Culcuilco Tres Zapotes Loma de la Coyotera
PACIFIC OCEAN
Tikal Pomona La Venta Palenque Yaxchilán
San José Mogote Monte Albán Gheo Shih
Tulum
Calakmul
El Mirador Nakbe Uaxactún San Bartolo San José BELIZE
Caracol
Siebal
HONDURAS GUATEMALA
San Lorenzo
Copán
EL SALVADOR NICARAGUA
Sites emphasized in text Sites mentioned in text Area of Mesoamerica
0
300 km
Modern international boundary
San José Mogote San Lorenzo El Mirador Monte Albán Teotihuacan Tikal Palenque Tula
site at its largest extent
_ _
Figure 8.2 Location of and timeline for sites in ancient Mesoamerica.
Chichén Itzá Tenochtitlán
episode of occupation
1500
1000
500
500
B.C./A.D.
Year
324
Chapter 8 Ancient Mesoamerica
1000
1500
Present
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 325
In Mesoamerica, the beginnings of sedentary village life mark the start of the Formative, or Preclassic, period. Studies at San José Mogote document this early stage of village living in highland Mesoamerica. The rapid transformation from simple village society to the construction of impressive ceremonial centers occurred precociously on Mexico’s Gulf Coast. San Lorenzo is the earliest of these lowland centers and is roughly contemporaneous with Poverty Point in northeastern Louisiana (see Chapter 7). El Mirador is unusually large and spectacular for the Maya Lowlands, considering its early date. Late in the Formative period, major urban centers were established in the highlands; hilltop Monte Albán in the Valley of Oaxaca and giant Teotihuacan in central Mexico illustrate this episode of development. By the beginning of the Classic period (ca. A.D. 200–300), Monte Albán and Teotihuacan grew significantly. The latter became one of the largest cities of the world at the time, even larger than Rome. Impressive centers, including Tikal and Palenque, also were built in the Maya Lowlands during the Classic era. Although the Classic Maya developed Mesoamerica’s most sophisticated writing system, none of their centers equaled the population size of Teotihuacan. Between A.D. 700 and 900, the Mesoamerican world underwent a sequence of upheavals and transitions that included the decline and depopulation of most extant centers. The succeeding Postclassic period was characterized by somewhat greater political fragmentation and fewer architecturally massive centers. Yet places like Tula, on Mesoamerica’s northern frontier, and Chichén Itzá, in the northern Yucatán, did rise to power for several centuries. The greatest exception to this Postclassic pattern was Tenochtitlán, which during the last years before Spanish conquest became the largest city in the history of prehispanic Mesoamerica. Although the rulers of Tenochtitlán established a tributary domain that stretched as far as highland Guatemala, they were defeated by armies led by Cortés in less than 2 years. In this chapter, our coverage emphasizes key large settlements that were the focal centers of prehispanic Mesoamerica (Figure 8.2). Yet it is crucial to understand that these centers often were linked through various mechanisms of exchange, interpersonal interaction, and shared systems of belief. Some of the longest-standing historical questions that remain for Mesoamericanist scholars to resolve concern the nature of the interactions between the Formative period Gulf Coast Olmec settlements and contemporaneous communities elsewhere in Mesoamerica and the kind of role that people from Teotihuacan, or influence from that central Mexican metropolis, may have had at Tikal and other Early Classic–period Maya centers. Another long-standing question concerns the nature and directionality of the connection between Tula and distant Chichén Itzá. Continued intellectual discoveries relating to these issues and others will only enhance the recognized richness of ancient Mesoamerican history. Unlike most states in the Eastern Hemisphere, Mesoamerican civilizations rose and flourished without beasts of burden, wheeled transportation, or metal tools. Yet during the 30 centuries that elapsed between the establishment of village farming communities and the Spanish conquest, this prehispanic world was the scene of highly developed statecraft, major urban centers, magnificent artisanship, spectacular architecture, and large swamp and lakeshore reclamation projects. Mesoamerica thus provides a physically diverse and scientifically important natural laboratory for joining historical and archaeological methods to unravel and interpret societal continuity and change. Today the prehispanic history of Mesoamerica receives too little attention in the history books of Europe and most of North America. Through this research and associated discoveries, we hope to bring greater attention to this region and the great contributions it has made to humankind’s heritage.
Thus they have come to tell it, thus they have come to record it in their narration, and for us they have painted it in their codices, the ancient men, the ancient women. Thus in the future never will it perish, never will it be forgotten, always we will treasure it, we, their children, their grandchildren, brothers, great-grandchildren, great-great-grandchildren, descendants, we who carry their blood and their color, we will tell it, we will pass it on to those who do not yet live, who are yet to be born, the children of the Mexicans, the children of the Tenochcans. —Fernando Alvarado Tezozomoc (Crónica Mexicáyotl, 1609)
_ Introduction Early State Development in Mesoamerica
325
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 326
Site
:
San José Mogote A 3500-year-old community in Mexico’s Southern Highlands
Figure 8.3 The low piedmont spur where San José Mogote was settled almost 3500 years ago. This natural rise was artificially enhanced to raise the top of the site’s principal mound, Mound 1, more than 15 m (50 ft) above the rest of the settlement.
The preceramic era in ancient Mesoamerica was long, and the pace of change was relatively slow (see “Guilá Naquitz Cave” and “Tehuacán,” Chapter 6, pp. 243 and 250). However, this period was marked by the origin of and increasing reliance on agriculture, as well as by steady but gradual increases in settlement size and occupational stability. The contrast with the subsequent episode of rapid change is pronounced. By the latter half of the third millennium B.C., the inhabitants of both coastal and highland settlements were making and using pottery. Within a few centuries, permanent villages were established across much of the Mesoamerican landscape.
_ _
326
Chapter 8 Ancient Mesoamerica
San José Mogote (SAN ho-ZAY mo-GO-tay), one of the more thoroughly studied of these early settlements, was positioned on a low spur of land abutting the flat valley bottom in the Valley of Oaxaca (wah-HA-kah) (Figure 8.3). This spot has been inhabited almost continuously to the present, and even today there is a small farming community at the site. First occupied before 1400 B.C., San José Mogote appears to be the oldest pottery-using village in the valley. A fairly complete archaeological record exists for the site between 1500 and 1150 B.C., when San José Mogote grew to become the largest—more than 2 ha (5 acres)—and most impor-
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
tant of the more than 25 villages distributed across the Valley of Oaxaca. At that time, San José Mogote was the only settlement containing several public buildings. Most of the early villages were situated on or near the valley bottom, the prime agricultural land. The inhabitants depended on maize, avocados, and other cultigens for subsistence, supplemented by wild plants from the piedmont and mountains; deer, cottontail rabbits, and other game were hunted. Small household units were equipped with braziers, earth ovens, and/or cooking hearths, as well as manos (grinding stones), and metates (ground stone basins), and blackened ceramic jars, for preparing and cooking food. Outside each household area were bell-shaped pits built for food storage and later used for trash disposal. Burials and other activity areas were located near the wattleand-daub dwellings. In total, each house and its surrounding features encompassed about 300 sq m (3200 sq ft). Clay figurines of humans and animals appear in association with rituals and burials in the first sedentary villages. Obsidian, a nonlocal stone from volcanic sources elsewhere in Mesoamerica, was found in varying quantities in each excavated residence. This evidence suggests that both trade and ritual were carried on at the level of the family during this period and that households were generally free of hierarchical control. Household autonomy began to diminish after 1150 B.C. Whereas most of the villages in Oaxaca remained small, San José Mogote continued to grow. Monumental architecture at the site was enhanced and enlarged. In the valley, certain craft activities were enacted only at San José Mogote. In one part of the town, magnetite was polished into mirrors that were traded as far away as the Gulf Coast, over 250 km (150 mi) distant. At the same time, exotic equipment for ceremonial activities—such items as turtle shell drums, conch shell trumpets, and stingray spines—were more abundant at San José Mogote than elsewhere.
Page 327
Obsidian also was more common there. Symbolic designs on pottery associated with the widespread connections of the Olmec Horizon are prevalent in Oaxaca from between 1150 and 800 B . C ., particularly at this most important town (see “The Olmec Horizon,” p. 335). After 700 B.C., differences between communities in the valley were even more pronounced. Major public buildings, incorporating adobe bricks and huge blocks of stone, were erected atop groups of earthen mounds. The valley’s largest architectural complex before 500 B.C. was built on a natural rise (Mound 1) that was artificially enhanced so that it rose 15 m (50 ft) above the rest of San José Mogote. As part of this complex, a flat stone was laid on the ground as a kind of threshold so that
Figure 8.4 Carved stone of sacrificed captive found near Middle Formative temple at San José Mogote.
brazier An open pan or ceramic ves-
sel used for holding hot coals. magnetite A black iron oxide that
can be polished to a lustrous surface. stingray spine Bony tail spines of
stingrays that were used in the past to draw blood in human autosacrificial rites.
_ Site San José Mogote
327
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 328
Preserved floor Preserved floor Structure
Burial (below floor)
Jades
Burial (under wall)
Tomb Tomb
Structure Mass of stone (possibly roof removed from tomb)
Patio
Structure
Preserved floor
0
danzante (Spanish) Dancer; a life-size
carving of a captive or a prisoner of war depicted in bas-relief on stone slabs at San José Mogote and Monte Albán, Oaxaca. Zapotec A Mesoamerican cultural
tradition generally associated with the Valley of Oaxaca and smaller adjacent valleys in central Oaxaca.
328
Step
Structure
2m
Figure 8.5 An elite residential complex, with tomb, that was constructed on top of earlier public buildings on Mound 1.
_ _
Preserved floor
N
anyone passing through a corridor between ceremonially important structures would step on it. The body of an awkwardly sprawled, naked individual with closed eyes, open mouth, and an open chest wound was carved on this slab (Figure 8.4). A similar theme, likely representing slain or sacrificed captives, appears slightly later with the carved stone danzante (dancer) figures at the site of Monte Albán (see p. 342). Slightly later in time at San José Mogote, the public buildings of Mound 1 were replaced by elite residential compounds, accessible only by the same stone stairway that had previously led to the public buildings (Figure 8.5). The importance of these residences is also suggested by the large storage areas associated with one of the compounds. A
Chapter 8 Ancient Mesoamerica
woman with three jade ornaments was buried in the same compound; the largest tomb at this time was situated in the central patio. Unlike the later palaces at Monte Albán, these elite compounds could have been built by the members of an extended family, without large labor gangs. Kent Flannery and Joyce Marcus, of the University of Michigan, who worked for two decades at San José Mogote, have concluded that the community’s highestranking family, however, did have sufficient power to direct the construction of a huge public building by corvée (unpaid) labor and eventually even to preempt that platform as a site for personal residence. Following the establishment of the city of Monte Albán (in 500 B.C.) on a hilltop less than 10 km (6 mi) to the south, San José Mogote became one of several second-tier communities that were part of the expanding political entity that was to become the Zapotec (ZAH-poh-tec) (Valley of Oaxaca) state. San José Mogote never regained the regional and interregional importance that it had before 500 B.C. Yet for the roughly 1000 years prior, beginning with the transition from mobile hunting and gathering to a sedentary way of life, San José Mogote was the central community in the Valley of Oaxaca. In Mesoamerica, cultural change was rapid following the shift to village life. At San José Mogote, a small farming hamlet expanded into a town with ceremonial and production functions encompassing more than 1000 people. These changes, which preceded the rise of Mesoamerica’s first urban civilization, are captured in the site’s archaeological record.
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 329
Concept
<
Nonresidential Architecture Clues to changes in the use of space and the nature of ritual
Anyone who has visited the archaeological sites of Mesoamerica remembers the massive pyramids, such as the Pyramid of the Sun and the Pyramid of the Moon at Teotihuacan, Temple I at Tikal, or the Castillo at Chichén Itzá. Nonresidential constructions, whether monumental, like the major Mesoamerican temple pyramids, or small, like Structure 6 at San José Mogote, can reveal significant information about the societies that built and used them. The size and complexity of construction can tell us about available labor and the power to organize and direct such projects. From the form or shape of public buildings, we can learn about the different kinds of communal activities that took place above the family level. Today such activities might include funerals, weddings, dances, sports events, parades, and political rallies. In prehispanic Mesoamerica, such activities included the ritual game played in formal ball courts (see “The Mesoamerican Ballgame,” p. 356), the burial of elite individuals in ornate tombs, and various ritual sacrifices enacted in and around the temples. The use of nonresidential space also tells something about accessibility to ritual activities. For example, an open plaza, or “danceground,” would have very different participants than the enclosed ritual space in the temples atop pyramid mounds. Changing patterns in the use of public space have been documented in a series of excavations in the Valley of Oaxaca by Kent Flannery and Joyce Marcus. The earliest known communal area was identified at the center of Gheo-Shih, a preceramic temporary campsite. Excavations revealed an arrangement of boulders in two parallel rows, each roughly 20 m (65 ft) long. The 7-m (23-ft)-wide space between the rows was swept clean and contained
N
0
5m
almost no artifacts, whereas outside the rock lines, artifacts were abundant (Figure 8.6). The space is thought to have been used for community rituals, perhaps a kind of danceground that was similar to those used by the Native Americans of Nevada during the seventeenth and eighteenth centuries A.D. A similar cleared feature was found in the early levels of occupation at San José Mogote, the oldest sedentary village in the Valley of Oaxaca. A new kind of construction appeared soon after the abandonment of this feature. A series of one-room structures, roughly 4 ⫻ 5 m in size (13 ⫻ 16 ft), were constructed of pine posts, walled with cane and clay, and plastered with lime (Figure 8.7). The first was built around 1350 B.C. and was continuously repaired and rebuilt on the same spot for centuries. These structures included a pit, filled with
Figure 8.6 The cleared area between the two rows of boulders, possibly used for communal rituals or dances, at Gheo-Shih.
Concept Nonresidential Architecture
_
329
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 330
Post molds Corner fragment
Plastered inner face
Lime-filled pit
Step or altar Outer (younger) stage
N
Inner (older) stage 0
1m
Post mold
Stones 2 layers high
Plastered outer face
Figure 8.7 Structure 6 is one example of an early public building at San José Mogote.
_ _
Figure 8.8 Remains of cut-stone architecture on a raised platform at San José Mogote.
330
Chapter 8 Ancient Mesoamerica
powdered lime, and a low bench along the interior wall. The lime may have been combined with narcotics, such as tobacco, and ritually smoked. Spanish chroniclers noted that the use of tobacco was prevalent among the Zapotec at the time of European contact. Although the specific rituals that took place in these early structures are unknown, only a limited number of the village’s inhabitants could have participated in them. Compared with the dancegrounds, access to the sacred world may have been somewhat more restricted at this time. Furthermore, the continual upkeep and construction suggest that increasing care and energy were devoted to public buildings and group ritual. Other elaborate public structures were erected at San José Mogote and elsewhere in the Valley of Oaxaca during the Formative period. These buildings often were placed on flattopped platforms of stone and adobe brick (Figure 8.8). By 600 B.C., this architecture was frequently faced with huge slabs of cut stone. Over time, the construction efforts in the Valley of Oaxaca reflected increased energy input and more restricted access to ceremonial activities. These two trends, along with the increasing diversity of Oaxacan public buildings, continued and intensified after 500 B . C . at the hilltop center of Monte Albán and elsewhere. This sequence of nonresidential building in the Valley of Oaxaca reflects changes in a society that began as a small, mobile population. At that time, the entire population of an open-air camp could participate in ceremonies, dances, or other rituals. With sedentism, larger populations, and the development of greater social, political, and economic differences within the Zapotec population, ceremonial activities became somewhat more exclusive and restrictive. The most elaborate architectural constructions are found at larger sites, such as San José Mogote and later Monte Albán. At these sites, access to these important civic-ceremonial buildings (and the activities and rituals within) was limited to a small, favored segment of the overall population.
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 331
Site
:
San Lorenzo and La Venta The roots of Mesoamerican civilization on the coastal plain of southern Veracruz
In 1860, a workman in southern Veracruz was clearing the dense tropical forest for a sugar plantation, when he noticed a huge, round object on the ground. He reported it to his foreman, who dispatched a crew to unearth it. What they had believed to be a large iron pot was actually the top of a colossal human head, carved of volcanic stone, standing 1.5 m (5 ft) tall (Figure 8.9). More than 50 years later, the place of discovery was recognized as the archaeological site of Tres Zapotes (TRACE zah-PO-tays). In the decades following this discovery, many other artifacts and artificial mounds were noted along the Gulf coastal plain in the states of Veracruz and Tabasco in Mexico. Yet archaeologists could not determine the age of these finds. The unique style of the monumental carved stone heads did not match the known archaeological traditions of the Maya, the Aztecs, or even pre-Aztec Teotihuacan. Few archaeologists suspected that these magnificent coastal materials were much earlier than those of the later, betterknown traditions. The age of the great basalt heads was not determined until the 1930s, when Matthew Stirling and his associates, of the Smithsonian Institution, began a 20-year project of mapping and excavation at the sites of San Lorenzo (sahn lore-REN-zoh), La Venta (la VEN-ta), and Tres Zapotes. These sites were located in dense, often swampy, tropical vegetation on broad river plains along the Gulf Coast. Stirling’s discoveries of carved basalt, extensive earthen mound groups, and caches of jade captured broad public and archaeological attention. Of great importance was the recovery of actual dates (around 31 B.C.) in the Mesoamerican calendar on carved stones at
Figure 8.9 A colossal human head from San Lorenzo. Called “The King,” this head is the tallest (2.85 m, or 9.35 ft) of the 10 heads that have been uncovered at the site. The largest carved heads weigh as much as 50 tons.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the economic foundations of the Olmec, see the Internet exercises on your online learning center.
Site San Lorenzo and La Venta
331
_
pri31057_ch08.pp320-389.qxd
10/15/09
10:59 AM
Page 332
Complex F 30
15 20 25
30 25 20 15
Complex E Complex I
Complex G Complex A 25
Complex C
30
Complex B
Central axis
Stirling “Acropolis”
25
20
Complex H
20
10
N
20
Complex D 15
Contour lines
0
100
200 m
Figure 8.10 The plan of La Venta. Complex A was the ceremonial core of the site.
Olmec The Aztec name for the late
_ _
prehispanic inhabitants of the Gulf Coast region of Mexico.
332
Tres Zapotes. These dates were older than any identified for the Maya and confirmed the antiquity of these materials, which are referred to as Olmec (see “The Olmec Horizon,” p. 335).
Chapter 8 Ancient Mesoamerica
Stirling’s pioneering work prompted more extensive studies at both La Venta and San Lorenzo. In the late 1950s, researchers from the University of California, Berkeley, used the radiocarbon method to date the principal Olmec occupation at La Venta to between 900 and 400 B.C. During that period, a series of clay and earthen pyramids were constructed on a low island in the middle of a swamp near the Tonalá River. The linear complex of structures, oriented 8° west of true north, covered an area more than 1.6 km (1 mi) long (Figure 8.10). One of the structures (in Complex C) was more than 30 m (100 ft) high. North of that high mound, a series of smaller platforms, plazas, and mounds were arranged symmetrically. Several platforms were built over mosaic pavements, each made of hundreds of blocks of greenstone. One mosaic depicts the symbolic mask or face of a jaguar. Offerings of greenstone, jade, and magnetite objects also were found in the excavations. The most spectacular of these, located under the center line of the architectural complex, included 16 greenstone human figures and 6 jade celts. These polished green objects (30 cm, or 12 in, tall) had been placed upright in a ceremonial arrangement (Figure 8.11). At its peak, La Venta was a very impressive center. The pyramids were built of clays of different colors, and the platforms were painted in various hues. Sculpted monuments, including four massive stone heads, were carved from volcanic basalt brought from about 100 km (60 mi) away. The work at La Venta made clear the antiquity of this early Gulf Coast occupation, as well as the labor, calendrical, and craft skills that may have been controlled by the Olmec rulers. These rulers were probably the individuals portrayed in the massive stone heads. A more recent project, conducted by Michael D. Coe, emeritus professor at Yale University, and Richard A. Diehl, emeritus professor at the University of Alabama, has provided significant information about the beginnings of the Olmec. At San Lorenzo, the Yale project mapped hundreds of monuments,
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 333
Figure 8.11 Offering 4 at La Venta. These 16 stone figures were arranged in a half circle under the architectural center line of Complex A.
mounds, and other features visible on the surface (Figure 8.12). These included over 20 lagunas, humanmade depressions that may have begun as borrow pits where dirt was removed for the construction of earthen mounds; later they were lined with waterproof bentonite (shalelike) blocks. The lagunas may have been used for ritual bathing. Part of the site was located on top of an artificially enhanced natural rise that stands 50 m (164 ft) above the surrounding countryside. Excavations revealed that San Lorenzo had a long and complex occupational history, reaching its peak between 1150 and 900 B.C., somewhat earlier than the major construction phase at La Venta. Significantly, the peak occupation at San Lorenzo was preceded by earlier village habitation. The first inhabitants settled only a small part of San Lorenzo, perhaps as early as 1500 B.C. By 1250 B.C., ceramic designs clearly related to the Olmec occupation were in use, and the villagers began to raise the natural mesa on which their houses were located. Deposits from 1250–1150 B.C. included basalt chips and lumps, discarded during monu-
ment carving, as well as Olmec blackand-white pottery. The Yale project established antecedents for the major occupation at San Lorenzo. This latter period was characterized by a major building episode that included massive mound and plaza construction, the carving of at least eight basalt heads and other monuments, and the layout of a drainage system to take water from the artificial lagunas. Also during this time, San Lorenzo developed ties with other localities throughout Mesoamerica. These connections are evident in the foreign goods found both at San Lorenzo and at sites in the highlands of Mexico, as well as in the widespread presence of the shared motifs of the Olmec Horizon. These designs probably had sacred or ritual importance throughout much of ancient Mesoamerica. For unknown reasons, building activities at San Lorenzo ceased after 900 B.C. The site was replaced, probably by La Venta, as the principal center of the Olmec. Michael Coe refers to the Olmec as Mesoamerica’s first civilization. Other archaeologists, such as the late William
laguna (Spanish) Lagoon.
Site San Lorenzo and La Venta
_
333
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 334
Figure 8.12 The central part of San Lorenzo, including the locations of mounds and lagunas.
North Court
Central Court
South Court
N
Hereditary rulership sanctified by divine descent is common to chiefdoms [societies with chiefs]. The colossal stone heads for which Early Formative Olmec culture is famous are probably portraits of the rulers of major Olmec centers. Many other of the stone monuments demonstrate in their iconography the divine descent and supernatural power of the rulers. A kinship tie between rulers at San Lorenzo and La Venta seems to be symbolized on one San Lorenzo altar, a relationship which would also imply hereditary rulership. —David C. Grove (1981)
pectoral A large ornament worn
_ _
across the chest, especially for defensive purposes.
334
0
50
100 m
Sanders, and Kent Flannery, suggest that San Lorenzo and La Venta were chiefly centers associated with social inequality and permanent leaders, but not with urbanism or state government. Clearly, the Olmec centers were occupied by elites as well as lower-status people. The former wore greenstone earplugs, hung iron-ore mirrors as pectorals, sat on sculpted thrones of basalt, and probably planned the ceremonial architectural precincts at the major centers. Lowerstatus people must have farmed and helped transport the construction materials of earth and basalt.
Chapter 8 Ancient Mesoamerica
Lagunas Mounds
Most of our information about the ancient Olmec comes from the ceremonial centers of a few key sites and the artifacts and sculptures associated with those high-status areas. Although in recent years archaeologists have started to collect more information about the size and layout of Olmec polities and conduct select excavations of villages and simple residences, it will be difficult to paint a more complete picture of this ancient social system until further research has been completed and fully published.
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 335
Concept
<
The Olmec Horizon Exchange and interaction in Formative Mesoamerica
On the Gulf Coast, the period between 1150 and 700 B.C. is distinguished by specific styles of monuments and art, including the massive basalt heads, large flat-topped mounds, rectangular carved stone altars (Figure 8.13), an abundance of greenstone, slant-eyed clay figurines, the were-jaguar figure (Figure 8.14), black-and-white pottery, and the repeated use of certain stylized designs on ceramic vessels (Figure 8.15). Although some of these elements, such as the huge stone heads, are known only from the Gulf Coast heartland, other traits, such as the paw-wing and crossed-bands motifs on pottery, are found both locally in the Gulf Coast and at early sedentary settlements across Mesoamerica. These widespread Olmec stylistic motifs have become recognized as elements of what archaeologists refer to as a horizon. The Olmec Horizon is not evenly distributed across Mesoamerica. Kent Flannery has noted that several Olmec designs are more common in the Valley of Oaxaca than along the Gulf Coast. Archaeologists have puzzled over the meaning of the Olmec Horizon. Some have suggested that it reflects Olmec conquests in the highlands or even an early empire, but that explanation seems doubtful, given the scale of the heartland Olmec sites and the wide distribution of the stylistic elements. The Gulf Coast Olmec were not powerful or populous enough to stake out and rule such far-reaching localities. Even the later and much larger Aztec state (with a large core population and a strong military) had great difficulty maintaining its broad tribute domain (the area from which it demanded tribute from local populations). Other archaeologists have proposed Olmec colonization of those areas where elements associated with
Figure 8.13 A side (top) and front (bottom) view of a carved stone altar from San Lorenzo. Although called altars, these carved stones are thought to have served as thrones for the rulers of Olmec sites.
were-jaguar A representation of a
supernatural figure that is half jaguar and half human. horizon A widely distributed set of
the horizon have been found. But that also seems unlikely; most of the participating regions had distinctive local traditions that existed before, and continued after, the adoption of Horizon
cultural traits and artifact assemblages whose distribution and chronology suggest they spread rapidly. A horizon is often composed of artifacts associated with a shared symbolic or ritual system.
Concept The Olmec Horizon
335
_
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 336
Figure 8.15 A stylized were-jaguar on a ceramic vessel.
Figure 8.14 A jade celt, the Kunz Axe, with unmistakable features of both a jaguar and a human child (actual size is 30 cm tall).
_ _
336
motifs. For example, in the highlands, local ceramic traditions generally were retained, incorporating the Olmec motifs into local ceramic complexes. Perhaps the most reasonable interpretation of the Olmec Horizon is that it represents an important symbolic component in a network of interaction that linked sites on the Gulf Coast with those elsewhere in Mesoamerica. Highland resources like obsidian, jade, and magnetite were found at the Gulf heartland centers, while mollusk and turtle shell, stingray spines, shark teeth, and Gulf Coast pottery were recovered in
Chapter 8 Ancient Mesoamerica
highland sites. In general, these exotic goods are found in ritual or highstatus contexts, as are the items carrying the Olmec symbols. The exotic goods and shared pottery designs may have served to enhance or affirm the social positions of certain individuals in both the Gulf Coast and the highlands. These social systems were characterized by increasing societal differentiation and the development of more formal leadership roles during this early village era. Clearly, the shared Olmec Horizon elements had a strong ritual aspect, with certain motifs serving as symbolic representations, or icons, for specific mythic creatures, deities, or kin-group totems, such as the were-jaguar. The sharing of rare materials and the ritual symbols also may have helped solidify communication and marriage alliances between groups of high-status individuals. The relative elaboration and abundance of the Olmec symbols in the heartland may signal the greater importance—and hence the expanded involvement in ritual, rule, and exchange—of the leaders at San Lorenzo compared with those at the smaller centers in other regions of Mesoamerica.
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 337
Site
:
El Mirador New perspectives on the beginnings of ancient Maya civilization Figure 8.16 A helicopter bringing in supplies to El Mirador, which is inaccessible by road.
The Maya civilization flourished in the tropical forests of the Petén lowlands of northern Guatemala between A.D. 250 and 900. Speculations about its beginnings traditionally have focused outside the region, partly because of the notion that the tropical forest was not suitable for large-scale agriculture. Some archaeologists saw the origin of the Maya as a direct extension of the Gulf Coast Olmec tradition, fostered by a belief that the Olmec were the “mother culture” of Mesoamerica, giving rise, like a pulsating beacon, to all later civilizations in the area. However, new interpretations of the Olmec Horizon have questioned this beacon model and have envisioned a network arrangement, in which certain concepts, knowledge, and symbols were shared by an emergent elite at a number of centers across Mesoamerica. With the more recent discovery of early preceramic and ceramic sites in the tropical forests of Guatemala and Belize, the origins of the Maya can now be traced locally. Outside influences or introductions from elsewhere in Mesoamerica are not essential to explain the
rise of this important civilization, although they may have contributed to the emergence of some later centers. Investigations at the sites of Nakbe (nok-BAY) and El Mirador (el mere-ah-DOR, “the lookout”) in the dense Petén forest of Guatemala have yielded new insights into the beginnings of Maya civilization. At Nakbe, some of the earliest large masonry public structures in the Maya Lowlands were built and occupied from 800–400 B.C. Later, by the last centuries of the first millennium B.C., Nakbe was surpassed by its neighbor, El Mirador. Surrounded by dense vegetation, more than 60 km (37 mi) from the nearest modern road or town, El Mirador is almost inaccessible today (Figure 8.16). Although there had been sporadic reports of these colossal ruins for more than 60 years, the first scientific visit took place in 1967, when Ian Graham made a preliminary map. Investigation of such a remote site is difficult. Research teams in the 1980s, including a project directed by Ray Matheny, of Brigham Young University, had to
_ Site El Mirador
337
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 338
Tigre Pyramid
Tigre Complex
Central Plaza Puleston Causeway Central Acropolis
Danta Complex Danta Pyramid
Monos Complex
Monos Pyramid
N West Group
0
1000 m
Figure 8.17 The architectural core of El Mirador, including the major pyramid complexes.
Tigre Pyramid
Structure 34
_ _
Figure 8.18 The Tigre Complex, the largest in the West Group at El Mirador. Viewed from the northeast in this reconstruction, the complex includes both the tall Tigre Pyramid and the much smaller Structure 34 in the southeast corner.
338
Chapter 8 Ancient Mesoamerica
East Group
pri31057_ch08.pp320-389.qxd
10/15/09
10:59 AM
transport food, mail, and supplies to the site by weekly mule trains; a landing strip had to be cleared to fly in equipment and personnel. In spite of El Mirador’s rugged setting, ceramic and architectural clues, as well as a few radiocarbon assays, date most of the known occupation at the site to between 300 B.C. and A.D. 250, before the Maya Classic period. El Mirador was huge—the largest known site of its time in the Maya area—and architecturally astounding. The site includes several hundred structures, some of which were larger than any built during the Maya Classic period. El Mirador is particularly important because there apparently was little building activity at the site during the later Classic period. Thus, unlike Tikal and most other Maya sites, the site’s Preclassic structures remain relatively undisturbed by later construction. El Mirador lies on a limestone plain dotted with broad, flat, clay-lined depressions, or bajos, which are filled
Page 339
with water for several months during the year. The site is not arranged according to a grid plan and appears to have been built in accordance with local topography. The structures at the site are distributed over 16 sq km (6 sq mi), with large, open spaces between buildings. The architectural core covers at least 2 sq km (0.8 sq mi) and is composed of a series of complexes: groups of plazas, platforms, buildings, causeways (raised roads), and other constructions (Figure 8.17). This core of the site is divided into two major building groups almost 2 km (1.2 mi) apart. The West Group is defined by a stone wall and ditch to the south and east and by steep escarpments to the north and west. Access to this area was controlled by a series of gates in the walls. These barriers may symbolically delimit the West Group as the sacred core of the site. Three features dominate the West Group: the Central Acropolis and the two huge platforms of the Tigre and Monos Complexes. The Tigre Complex,
Figure 8.19 Archaeologists uncovering a stucco mask on Structure 34 in the Tigre Complex.
bajo (Spanish) A broad, flat, clay-lined depression in the Maya Lowlands that fills with water during the rainy season.
Site El Mirador
339
_
pri31057_ch08.pp320-389.qxd
10/1/09
E Group An arrangement of buildings
designed to mark the position of the rising sun during important solar events, such as equinoxes and solstices, in Mesoamerica. stucco A type of plaster, often made
of lime, used for decoration. facade The face, or front, of a building.
8:11 AM
with an enormous pyramid 55 m (180 ft) high, apparently was constructed primarily during the last centuries B.C. (Figure 8.18). This complex forms what Mayanists refer to as an E Group. Structure 34, a small temple included within the Tigre Complex, had large stucco sculptures on its facade. The sculpted stucco masks, which incorporate both human and jaguar features, are very similar in style and context to those at other sites in the Maya Lowlands (Figure 8.19). The messages transmitted through these symbols may convey divine sanction for the right to rule. Although the huge Monos Complex has still not been fully explored, the Central Acropolis has been mapped and appears to be the nucleus of the site. Measuring 330 m (1080 ft) long by roughly 100 m (325 ft) wide, this structure supports a number of buildings, including several possible palaces or high-status residences. A causeway leads from the east gate of the West Group to the East Group. This group is known largely for the Danta Complex, a massive series of sculpted terraces and platforms, built on a natural elevation that was leveled and modified to accommodate
_ _
340
Page 340
Chapter 8 Ancient Mesoamerica
the monumental construction. From the top of the highest building in the Danta Complex, one can see above the forest canopy for 40 km (25 mi) in all directions. El Mirador could not have existed in a vacuum. Food and labor for construction almost certainly had to come to the site from the surrounding countryside. The movement of people and materials followed a series of six long causeways built like the spokes of a wheel from the site’s hub. These impressive causeways would have been especially useful during the summer rainy season, when the bajos and low ground turn to heavy and sticky mud, becoming impassable. Much of El Mirador is still unexplored, hidden under thick tropical vegetation. Some questions remain about the actual size of the site and how much of it was built before A.D. 250, the start of the Maya Classic period. Yet there seems little question that El Mirador and other early Maya sites such as Nakbe have revolutionized archaeological views of both the Preclassic in the Petén lowlands and the origins of Maya civilization.
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 341
Concept
<
Carved Stones and Early Writing Conveying important prehispanic Mesoamerican messages
From the carved heads and thrones of the Olmec to the circular calendar stone of the Aztecs, stone was a primary medium of prehispanic Mesoamerican art. At almost every major site, archaeologists have recovered sculpted stone monuments. Such stones were used as architectural elements, for the facades of earthen or rubblefilled structures, or as freestanding monuments, such as the thrones and giant heads of the Olmec or the upright stone shafts, or stelae, from later periods. Such stelae often glorified the ancestries and accomplishments of specific rulers or elites. Freestanding stelae are particularly important because they contain examples of the earliest Mesoamerican writing. The first carved stones with hieroglyphs, found in the Valley of Oaxaca and at the Gulf Coast site of La Venta, date to between 600 and 400 B.C. The inscription from Oaxaca contains two glyphs that indicate “1 Earthquake.” This inscription represents a date in the Mesoamerican calendar (see “Writing and Calendars,” p. 369). It is the first documented use of that sacred system. These glyphs were carved between the feet of an individual who was probably a captive or sacrificial victim. In Mesoamerica, individuals often were named by their birth date, and Joyce Marcus has suggested that 1 Earthquake may be the name of the prisoner depicted. The La Venta stone, the so-called Ambassador Monument, includes three glyphs, a central figure, and an isolated footprint (Figure 8.20). Later in prehispanic Mesoamerica, the footprint was a conventional symbol for travel or journey. Writing, calendrics, and the use of stelae were developed to their greatest extent by the Maya of the Petén lowlands. The largest number of texts
pertain to this region and date to between A.D. 250 and 909. Yet recently, in 2001, the oldest Maya writing, dating to around 100 B.C., was found at San Bartolo in the Petén lowlands of Guatemala. William Saturno, of Boston University, was examining a looter’s tunnel at the site when he discovered painted murals on interior walls of a Late Preclassic building. The murals were accompanied by short hieroglyphic texts. Yet because the glyphs are much older than previously known Petén writing, most have not yet been deciphered. These texts do not include any Long Count dates. Some of the earliest stone monuments do include calendar dates in the Long Count, a cycle with a fixed zero date that can record large blocks of time (see “Writing and Calendars,” p. 369). These monuments have been found at sites in the Maya Highlands and along the Pacific Coast. The earliest Long Count dates fall in the last century B.C. and the first century A.D., years before the first Petén stela, which was erected during the third century A.D. (see “Tikal’s Monument Record,” p. 361). These carved stones must have carried great importance for prehispanic Mesoamericans. Significant labor was devoted to both their construction and, at times, their destruction. At Olmec and later Maya sites, there is evidence for the intentional mutilation and defacement of these stone monuments, possibly to diminish the importance of the individuals who were represented (perhaps after their death), just as statues may be toppled in a revolution today. Yet, despite the destruction, the carved stones are durable and provide an important record for archaeologists. Unfortunately, the many later Mesoamerican documents on bark paper and deerskin did not fare as well.
Figure 8.20 Inscriptions on the Ambassador Monument from La Venta.
stela (plural stelae) (Latin) An erect
stone monument that is often carved. hieroglyph Originally, the pictographic script of ancient Egypt; any depictive, art-related system of writing, such as that of Mesoamerica. Also may refer to an individual symbol. glyph (Greek) A carving; a drawn
symbol in a writing system that may stand for a syllable, a sound, an idea, a word, or a combination of these.
_ Concept Carved Stones and Early Writing
341
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 342
Site
:
Monte Albán A hilltop city in the Valley of Oaxaca Building over Tomb 104
Figure 8.21 The Main Plaza at Monte Albán.
North Platform
Sunken patio Ball court
Danzantes Gallery
Tunnel
Building J
South N Platform 0
100 m
The Valley of Oaxaca is divided into three major arms by the Atoyac River and its tributary, the Salado. The northern arm was the primary focus of early settlement between 1500 and 500 B.C., particularly around San José Mogote. Although several large villages were established in the valley’s southern and eastern arms after 700 B.C., San José Mogote continued as the region’s largest and architecturally most im-
_ _
342
Chapter 8 Ancient Mesoamerica
pressive locality. Around 500 B.C., the pattern of settlement shifted, following the establishment of the hilltop center of Monte Albán at the hub of the valley’s three arms. The monumental hilltop ruins of Monte Albán (MOAN-tay al-BAHN) have attracted explorers, antiquarians, and archaeologists for well over a century. Systematic excavations were begun at the site in 1931. These studies were carried out over 18 field seasons by Mexican archaeologist Alfonso Caso and his colleagues, Ignacio Bernal and Jorge Acosta. Their research involved the opening of more than 170 tombs, a series of stratigraphic excavations, and the clearing and reconstruction of many buildings. Much of their work focused on the Main Plaza at Monte Albán, an impressive concentration of architecture erected on the artificially flattened summit of the hill (Figure 8.21). In 1971, Richard Blanton, of Purdue University, initiated a survey to map the ruins of Monte Albán and to define changes over time in the size and organization of the site (Figure 8.22). From its foundation, Monte Albán was unique in the Valley of Oaxaca. Whereas earlier settlements were on or next to the valley bottom, Monte Albán was situated above and at some distance from good farmland. Blanton’s survey indicates that the city also grew very rapidly to a size of more than 1 sq km (0.4 sq mi). During the last centuries B.C., it had an estimated population of 15,000 people and, after its foundation, quickly surpassed San José Mogote in size and architectural elaboration. Soon after the establishment of Monte Albán, major building activities ceased at San José Mogote for several centuries. During the phase known as Monte Albán I, 500–200 B.C., more than 300 stone monuments were carved and dis-
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 343
Figure 8.22 A map of structures and residential terraces in the core area of Monte Albán. Although the most monumental buildings at the site surround the Main Plaza, groups of smaller buildings are dispersed through the residential sectors of the city.
N
Ancient roads Defensive walls Mounded buildings Residential terraces
0
.25
.5 km
played in one of Monte Albán’s first public buildings (Figure 8.23). Although the figures on these stones have been named the danzantes (dancers), most of them are similar to those on an earlier monument with a military theme at San José Mogote. Naked and sprawled in grotesque, rubbery positions, with closed eyes and blood running from one or more wounds, the danzantes probably represent captives or prisoners of war. Although the specific meaning of these carved stones remains uncertain, their date corresponds in
time to the consolidation of power at Monte Albán and the political unification of the Valley of Oaxaca. Many other changes took place in the Valley of Oaxaca with the rapid growth of Monte Albán. Monumental construction increased. Greater concern with the dead (and with their ties to the living) was reflected in the construction of elaborate subterranean tombs. Domestic architecture evolved from simple wattle-and-daub structures to more permanent adobe constructions. Greater regional similarities
_ Site Monte Albán
343
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 344
Figure 8.23 Danzantes displayed in the Danzantes Gallery, an early public building on the Main Plaza at Monte Albán.
Figure 8.24 The arrowhead-shaped Building J at Monte Albán was erected in the center of the Main Plaza.
in pottery styles also were noted. New ceramic forms, including the comal and the incensario, suggest changes in both cooking technology and ritual activities at that tme. The rapid growth in the number and size of settlements in the Valley of Oaxaca is further indication that this was a period of dynamic transition.
_ _
344
Chapter 8 Ancient Mesoamerica
During the second phase of Monte Albán (Monte Albán II), 200 B.C.– A.D. 200, the site expanded its political, economic, and military influence beyond the Valley of Oaxaca. At the same time, a 3-km (2-mi)-long defensive wall was erected at the site. An unusual building with an arrowhead-shaped ground plan (Building J) was erected
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
in the Main Plaza (Figure 8.24). Set into the walls of this structure were more than 40 carved stone panels. Each panel included a set of glyphs with a place name, probably identifying a location, along with an upsidedown human head, which likely signified military defeat (Figure 8.25). Caso identified these slabs as the documentation of Monte Albán’s conquests, and Joyce Marcus has suggested identifications for several distant localities represented on the stone slabs. In one of these identified localities, the Cuicatlán Cañada (kwee-caht-LAN cahn-YA-da), 80 km (50 mi) northeast of Monte Albán, Charles Spencer, of the American Museum of Natural History, and Elsa Redmond, of Columbia University, have found strong indications of Monte Albán conquest during Monte Albán II. A local community, the site of Loma de la Coyotera (LOH-mah day la coy-yo-TARE-ra), was sacked and burned, with the skulls of 61 of its inhabitants mounted on a rack (Figure 8.26). This skull rack, what the Aztecs called a tzompantli, was found to be contemporaneous with Monte Albán II. Tzompantli erection was generally associated with military conquest during Aztec times and probably served similar functions in earlier phases as well. Spencer and Redmond also found a major transition in the settlement pattern of the Cuicatlán Cañada, including the construction of a fortress that was placed to control traffic through the Cañada and between the Valleys of Oaxaca and Tehuacán. Monte Albán’s dominance over the Valley of Oaxaca continued during the third phase of its history (Monte Albán III), A.D. 200–800. However, during this phase, its influence outside the valley diminished, perhaps because of the growing importance of Teotihuacan, the great urban center in the Valley of Mexico (see “Teotihuacan,” p. 349). In the Cuicatlán Cañada, for example, the Monte Albán connection waned during this period, and settlement patterns and ceramic findings point to the increasing influence of Teotihuacan.
Page 345
Late in Monte Albán III, Monte Albán itself reached its greatest physical size, with a population of 25,000–30,000 people. Sometime after A.D. 700, major construction at Monte Albán ceased, and the city began to decline in size and importance. Although a remnant population occupied the hilltop for much of the rest of the prehispanic era, the city never again achieved its former glory. Toward the end of prehispanic times, a powerful ruler was buried in a tomb originally built during Monte Albán III at Monte Albán. This subsurface feature (Tomb 7) contains ancient Mesoamerica’s greatest treasure, a tomb stuffed with gold, carved jaguar bones, shell, turquoise, jet, crystal, and numerous other exotic items. Monte Albán was quite different in plan, size, and function from the contemporaneous Mesoamerican centers of Teotihuacan or Tikal or the later Aztec city of Tenochtitlán. Monte Albán’s hilltop location is not typical of Mesoamerican sites as a whole. With its monumental Main Plaza, dwarfing the architecture found at contemporary sites in the valley, Monte Albán’s politico-religious functions always surpassed its regional commercial role. In
Figure 8.25 A place glyph, with an upside-down head, on one of the conquest slabs erected on Building J at Monte Albán.
Figure 8.26 An artist’s reconstruction of the skull rack (tzompantli) at Loma de la Coyotera.
_ Site Monte Albán
345
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Figure 8.27 Monte Albán, as seen from the air, towering over the surrounding countryside. The Main Plaza occupies the artificially flattened summit of the hill, with thousands of residential terraces carved into the descending slopes.
_ _
The growth of the early civilization represents one of anthropological archaeology’s most challenging but unmanageable problems. It is unmanageable because the complexity and variety can overwhelm. Given this, there are basically two strategies that researchers can adopt. One is to ignore the complexity and try to find one factor or only a few that cut right to the core of things and offer a maximum of explanatory power with a minimum of bother. The other is to try to match the complexity of the reality with complexity and richness in theory and method. The latter probably dictates that the researcher be grounded in ethnology, demography, economics, political organization, home economics, human biology, ecology, and probably several other disciplines as well. —Richard E. Blanton (1983)
346
Page 346
general, archaeological indicators for high-volume craft production are surprisingly few at the hilltop city, and no central market area has yet been securely identified, although a broad plaza below the Main Plaza may have served as a marketplace. The site was not laid out according to a grid plan. Few major roads were constructed, and access to the Main Plaza was restricted to several narrow entryways. The bulk of the population lived on the more than 2000 residential terraces carved out of the hillside beneath the Main Plaza and on adjacent hills (Figure 8.27). Held in place by stone retaining walls, each terrace contained no more than a few houses. Based on recent research at Monte Albán and contemporaneous settlements in the Valley of Oaxaca, these houses generally included a series of rectangular rooms around a limeplastered patio. The production of craft goods for exchange was carried out in these residential complexes. By 200 B.C., Monte Albán was the capital of a Zapotec state whose limits
Chapter 8 Ancient Mesoamerica
stretched beyond the Valley of Oaxaca. Although archaeologists have yet to unravel all the factors responsible for either the foundation of this hilltop center or the rise of the Zapotec state, these changes seem in part tied to the increasing role of militarism in Oaxaca. The danzantes, the conquest slabs, Monte Albán’s hilltop location and its defensive wall—all point to the importance that warfare may have had. Yet the emergence of Monte Albán also coincided with marked regional population growth, as well as changes in the local systems of exchange and production. In regions bordering the Valley of Oaxaca, marked shifts in settlement size and nucleation also were taking place. All these factors could have had key roles in the rise of Monte Albán and associated changes.
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 347
Concept
<
Settlement Pattern Surveys Examining ancient demography and the distribution of sites on regional landscapes
Traditionally, archaeologists have depended on excavation as the primary means for obtaining information. Paradoxically, however, when archaeologists excavate sites, they also destroy the buried deposits of material. Unlike physicists or chemists, who can repeat experiments numerous times under the same conditions, archaeologists can dig a particular area only once. Excavations, therefore, proceed slowly and carefully, to record as much detailed information as possible. Because of limited time and money, we are lucky to investigate more than a small part of any one site or more than a few sites in a given area. Sociologists would be uncomfortable describing a modern city from only a few interviews and often supplement detailed accounts with broad urban surveys or censuses. For similar reasons, geologists and geographers often combine very specific information gathered from soil probes or individual field studies with the larger-scale perspective provided by satellite imagery. Over the past 50 years, archaeologists have developed a technique, called systematic settlement pattern survey, for studying areas larger than one or a few sites. Archaeologists have for a long time used surface survey procedures to locate sites based on the presence of artifacts and architectural remains on the ground. They frequently use unsystematic surveys, or reconnaissances, to find sites to excavate. Over the past several decades, archaeologists have refined regional archaeological survey procedures. In systematic surveys, large blocks of land are thoroughly and carefully walked over to locate surface evidence of past occupations. These
occupations, which are dated by the artifact scatters associated with them, are placed on aerial photographs or topographic maps. The entire distribution of sites known for each particular period represents that era’s settlement pattern. The first systematic settlement pattern survey was carried out in the Virú Valley on the north coast of Peru soon after World War II by the late Gordon Willey. Willey designed this study to determine the geographic and chronological position of sites, to outline the developmental history of settlement, and to reconstruct cultural institutions as far as they were reflected in the settlement data. To achieve these ends, Willey and his colleagues walked over the valley, mapping a total of 315 archaeological sites. These sites were dated and classified according to estimated function, determined primarily by site size, location, and the kind of architecture visible on the surface. Pieces of pottery on the ground surface were used to date the sites. Through this extensive surface survey, Willey obtained a regional overview of changing patterns of settlement that he could supplement with detailed excavations at a few sites. Since Willey’s pioneering study, systematic settlement pattern surveys have been undertaken in many areas, including the highlands of Oaxaca and central Mexico, the rivers and foothills of Southwest Asia, the mountain valleys of the Andes, and the deserts and plateaus of the southwestern United States. Surveys are most successful where artifacts are readily visible on the ground. Such conditions are found frequently in semiarid climates, where vegetation is limited and soil formation is slow.
Settlement pattern surveys provide very different kinds of information than excavations. What can the findings of systematic surveys tell us about ancient societies? Why are these studies important for our interpretations of the past?
y
_ Concept Settlement Pattern Surveys
347
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 348
Figure 8.28 Gary Feinman (left) and Stephen Kowalewski (right) surveying the archaeological site of Jalieza, the Valley of Oaxaca, in 1977.
Archaeological surveys provide a variety of information about the past. The number and sizes of sites enable the archaeologist to estimate ancient population size and changes in demography over time. Archaeologists also can examine the distribution of sites on the landscape in relation to different kinds of soil, topographic features, and resources. Like household interviews and censuses, soil probes and satellite images, excavations and surveys provide complementary insights. Excavations yield detailed, specific, more precise observations at a small scale. Surveys provide broader, less fine-grained information at a larger scale. When used alone, each method can provide misleading results because of small sample size (excavation) and the reliance on surface materials no longer in their primary context (survey). The most _ _
348
Chapter 8 Ancient Mesoamerica
complete descriptions are achieved when both are used. The human past is most discoverable when we have perspectives or observations at a range of scales, from the activity area and domestic contexts through the neighborhood, site, region, and even geographic domains larger than regions. Now more than 30 years after the advent of systematic regional settlement pattern study in the Valley of Oaxaca, survey crews have covered a contiguous area of more than 8000 sq km (3100 sq mi) in this highland region (Figure 8.28). This macroregional coverage is giving researchers the chance to understand the patterns of relationships between the Valley of Oaxaca and adjacent areas over millennia.
pri31057_ch08.pp320-389.qxd
10/1/09
8:11 AM
Page 349
Site
:
Teotihuacan One of the world’s largest cities in A.D. 500
Lake Zumpango
Cerro Gordo
Lake Xaltocán
Teotihuacan
2800
Cerro Patlachique
280
0
00
34
Basin of Mexico Cuicuilco
Lake Xochimilco
30 00
Lake Texcoco
3000
Lake Chalco
30
00
0
40 00 4
44
Iztaccihuatl 3000
3000
2800 2000
Popocatepetl
1600
4400
Today Mexico City, with over 19 million inhabitants, covers most of a huge geological basin in the Central Highlands of Mexico. Before the enormous growth of Mexico City, the Basin of Mexico contained the largest expanse of flat, agricultural land in all of highland Mesoamerica. The basin lies at an elevation of 2100 m (7000 ft) and is ringed by a series of largely dormant, snow-capped volcanoes forming a palisade around its perimeter (Figure 8.29). The outlet for the basin was dammed by a number of volcanic eruptions several million years ago, forming a series of lakes. The floor of the basin originally held five shallow lakes, brackish in the north and fresher to the south. These lakes, of particular importance for the Aztec civilization, were drained by the Spanish during the Colonial period, when Mexico was ruled by the Spanish crown (A.D. 1521–1821), and only small remnants of them remain. The Basin of Mexico was the center of at least two major prehispanic civilizations. Systematic archaeological surveys of more than half the floor of the basin, made during the past 50 years, have shown that the first sedentary villages were concentrated in the southern half. This area receives more rainfall than the northern portion of the basin and is better for farming. Before 500 B.C., occupation was established at a site called Cuicuilco (kwee-KWIL-co) in the southwestern part of the basin. Relatively little is known archaeologically about this site. It was buried by at least two prehispanic lava flows and then covered by the urban sprawl of modern Mexico City. The tops of only a few pyramids rise above the mantle of volcanic stone. Yet the scale of Cuicuilco’s distinctive structures and the large area of its occupation suggest that the site was the basin’s first major town.
3000
2800
N 2000
1600 0
8
16 km
The volcanic eruptions may have helped end Cuicuilco’s preeminence and may have contributed to a shift in the balance of power in the basin. By the end of the second century B.C., the site of Teotihuacan (tay-o-tee-WA-kahn) had been established in the northeastern corner of the basin, about 20 km (12 mi) from the lakeshore. Situated in low hills adjacent to a series of natural springs, Teotihuacan grew rapidly, becoming one of the region’s two primary centers, along with Cuicuilco.
Figure 8.29 The Basin of Mexico, showing the prehispanic lakes, the volcanoes, and the major early centers of Cuicuilco and Teotihuacan. The lakes are now covered by modern Mexico City.
_ Site Teotihuacan
349
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 350
Pyramid of the Moon N
Pyramid of the Sun
West Avenue
Great Compound
Avenue of the Dead
San Juan River
Figure 8.30 The central core of Teotihuacan, showing the principal avenue, the Avenue of the Dead, and the major buildings.
_ _
350
Ciudadela Temple of the Feathered Serpent
East Avenue
Based on his urban survey of Teotihuacan, René Millon, emeritus professor at the University of Rochester, estimates that the site may have reached 6–8 sq km (2.5–3.0 sq mi) in area, with a population of at least 20,000 by this time. The population of the basin became increasingly nucleated during this period, particularly around the major centers. At the same time, areas of little habitation formed between the population clusters, suggesting the emergence of buffer zones between hostile groups.
Chapter 8 Ancient Mesoamerica
By the first century A.D., Teotihuacan had become the primary settlement in the Basin of Mexico. While the urban center sprawled to almost 20 sq km (7.5 sq mi), encompassing roughly 90,000 people, other centers in the basin declined in size and importance. In fact, the next largest communities had populations of less than 5,000 people and comparatively little nonresidential construction. This tremendous nucleation of population and power continued over subsequent centuries, signaling the dramatic political and economic reorganization that took place during the hegemony (predominance) of Teotihuacan. Several hypotheses have been advanced to account for the rapid growth and development of this great city. We will probably never know just how closely the growth of Teotihuacan was related to the natural disaster that buried a portion of the contemporary center of Cuicuilco in the last centuries B.C. Alternatively, William Sanders has emphasized the importance of Teotihuacan’s location, positioned to tap springs for a series of irrigation canals. For decades, Sanders and other archaeologists have been influenced by the writings of Karl Wittfogel, who emphasized the managerial costs of water control and the possible role such management had in stimulating political change. Although canals were built to harness the water from these springs, the specific timing of this construction has not been verified. Sanders and others have suggested that canal construction and agricultural intensification would have been required to feed the burgeoning population of this ancient city and the basin as a whole. Noting the abundance of obsidian at Teotihuacan, Michael Spence, of the University of Western Ontario, has suggested that the manufacture of obsidian artifacts may have played a role in the growth of the city, something like the relationship between the automobile and Detroit. In the ancient Mesoamerican world, obsidian provided the sharpest cutting edge available. It was used economically for knives and other tools, ritually to extract blood in
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
autosacrifice (the drawing of blood from the fleshy parts of one’s own body), ornamentally as part of masks and necklaces, and militarily for weapons. Raw obsidian nodules can be taken from the San Juan River, which cuts directly through Teotihuacan, as well as from a large source of gray obsidian only 16 km (10 mi) away. Teotihuacan also is close to a source of green obsidian at Pachuca, roughly 50 km (30 mi) to the northeast. Pachuca obsidian, present in great quantity at Teotihuacan, was valued for both workability and the symbolic importance of its color. Whereas Spence has emphasized the role of obsidian production and export in Teotihuacan’s rise, John Clark, of Brigham Young University, has suggested that most of the surface obsidian debris could reflect production for local use. Regardless of its eventual destination, there is little doubt that the quantity of obsidian at Teotihuacan far exceeds what is typical in Mesoamerican sites. Millon has stressed the importance of ideological, as well as economic, factors in the developmental history of Teotihuacan. It was a sacred city with more temples than any other prehispanic Mesoamerican site, before or after. At its maximum size, between the fifth and seventh centuries A.D., Teotihuacan covered an enormous area of 22 sq km (8.5 sq mi), with an estimated population of at least 125,000 people. The city was planned and laid out along a rectilinear network of roads and paths. The major north-south axis, the Avenue of the Dead, intersected with lesser East Avenue and West Avenue (Figure 8.30). Together these avenues divided the city into quadrants. At their intersection were the distinctive Ciudadela (citadel) and the Great Compound, which formed an enormous architectural unit. The Great Compound is thought by some scholars to have been the city’s central marketplace. The Ciudadela was a huge political and religious precinct, enclosing a sunken courtyard and the imposing Temple of the Feathered Serpent. The Ciudadela, together with all the buildings along the Avenue of the Dead, was constructed with cut-stone
Page 351
(a) Talud Stone slab Tablero
Lime plaster Volcanic stone set in clay (b)
Figure 8.31 Talud-tablero architecture: (a) a typical structure; (b) a cross section showing the elements of construction.
facing known as talud-tablero (Figure 8.31). This style, composed of framed panels (tableros) and sloping basal elements (taluds), is recognized as a symbol of Teotihuacan temples. Its presence on the entire 2 km (1.25 mi) of structures from the Ciudadela to the Pyramid of the Moon would seem to denote the sacred aspect of the central city. Saburo Sugiyama, of Aichi Prefectural University, Japan, and Rubén Cabrera, of Mexico’s National Institute of Anthropology and History, have recently excavated burial chambers under the Pyramid of the Moon and the Temple of the Feathered Serpent in the Ciudadela that contain the remains of royal retainers and sacrificial victims. However, no royal tombs have been uncovered. The elaborate offerings found with the burials, such as imitation human teeth carved from shell, are quite different from artifacts found in residential areas.
It has mainly been the humanists who have studied the informational aspects of complex societies—art, religion, ritual, writing systems, and so on. The “ecologists” have largely contented themselves with studying exchanges of matter and energy—the “technoenvironmental” factors. . . . To read what the “ecologists” write, one would often think that civilized peoples only ate, excreted, and reproduced; to read what the humanists write, one would think civilizations were above all three, and devoted all their energy to the arts. . . . I will argue that humanists must cease thinking that ecology “dehumanizes” history, and ecologists must cease to regard art, religion, and ideology as mere “epiphenomena” without causal significance. —Kent V. Flannery (1972a)
talud-tablero (Spanish) An architec-
tural style characteristic of Teotihuacan during the Classic period, in which recessed rectangular panels (the tablero) are separated by sloping aprons (the talud).
Site Teotihuacan
351
_
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Figure 8.32 The Pyramid of the Sun at Teotihuacan. The monumental pyramids at the site are among the largest structures ever built in the ancient Americas.
y
Some of the largest structures ever built in the ancient Americas were constructed at Teotihuacan. When and why are giant pyramids and monumental structures built?
_ _
352
Page 352
The largest of the city’s more than 5000 known structures, the Pyramid of the Sun and the Pyramid of the Moon dominate the surrounding landscape (Figures 8.32 and 8.33). The monumental Pyramid of the Sun was the largest structure ever built during a single construction episode in the ancient Americas. The structure stands 64 m (212 ft) high and measures roughly 213 m (700 ft) on a side. It contains 1 million cu m of fill (35 million cu ft), which was carried to the pyramid in basketloads. The cut-stone exteriors of these massive structures (as well as most other platforms at the site) are believed to have been faced with a thick, white plaster and painted red to enhance their visibility. Hidden below the Pyramid of the Sun lies a cave. It runs 100 m (330 ft) from its mouth near the base of the pyramid stairway to a spot close to the center of the pyramid. The people of Teotihuacan must have used the cave, because ritual items were found within it, and in certain spots, the walls of the
Chapter 8 Ancient Mesoamerica
cave were reroofed. Caves were sacred in Mesoamerican religion, associated with the creation of the sun and the moon. Throughout Mesoamerica, important ritual activities were enacted in caves. Archaeologist Linda Manzanilla, of the National Autonomous University of Mexico (UNAM), has noted that the cave under the Pyramid of the Sun was just one of many at the site, and she has suggested that these tunnels and caverns were the source of virtually all the volcanic stone used to build Teotihuacan. The residential pattern for the earliest years of Teotihuacan is not well known. However, a series of distinctive multihousehold residential units dating to the third century A.D. can be recognized. Some 2200 of these wellplanned, single-story apartment compounds were eventually built. During the site’s later history, these compounds were the principal kind of residential structure (Figure 8.34). The interiors of the compounds were divided into different apartments, each with rooms,
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 353
Figure 8.33 A view of the Pyramid of the Moon and the north end of the Avenue of the Dead from the top of the Pyramid of the Sun at Teotihuacan.
Temple
Temple
Temple
N
0
10
20 m
Figure 8.34 The plan of an apartment compound at Teotihuacan. The areas in color are roofed; the uncolored areas are open plazas or temples.
_ Site Teotihuacan
353
pri31057_ch08.pp320-389.qxd
10/1/09
Figure 8.35 Drawing of processional figures in mural at Teotihuacan.
_ _
patrilocal Describing a residence pattern in which married couples live with or near the husband’s family.
354
8:12 AM
Page 354
patios, and passageways. The exteriors were surrounded by tall, windowless stone walls of cement and plaster. The compounds varied significantly in size. Millon estimates that the average structure housed about 60 people, while the larger ones held 100 or more. From size, architectural differences, and the kinds of artifacts and wastes found at the compounds, it is clear that the occupants of these units varied in socioeconomic position. The groups that lived in the compounds were enduring. Many of the structures were rebuilt several times over several centuries, with little change in plan. The specific relationships that linked domestic units in a compound remain unknown. Spence’s preliminary study of the skeletal materials led him to speculate that the males within one compound had fairly close biological ties, whereas the females did not. Such a pattern suggests a lineage or extendedfamily residence pattern that was patrilocal, in which females moved in with the family of the groom after marriage. Spence also has suggested that individuals within apartment compounds may have shared certain economic skills, such as working obsidian. Each compound had at least one temple or shrine, suggesting joint participation in ritual by the residents.
Chapter 8 Ancient Mesoamerica
Certain neighborhoods at Teotihuacan were associated with foreign residents. The compound occupied by people from Oaxaca included a tomb with an inscribed stela bearing a glyph and number in the distinctive Oaxacan style. One ceramic urn came directly from the Valley of Oaxaca; other pottery vessels were made from local basin clays in the Oaxacan style. At least some of the foreign residents are assumed to have been merchants or traders. In another neighborhood of Teotihuacan, the houses were built of adobe following the style of the Mesoamerican Gulf Coast. Teotihuacan was involved in trade relationships that extended as far as the Gulf Coast of Mexico, the Maya Lowlands, the Guatemalan highlands, and the deserts of northern Mexico. On Monte Albán’s Main Plaza, carved stone monuments depict an important, yet apparently peaceful, meeting between an emissary wearing a costume from Teotihuacan and a Zapotec lord. Figures bearing symbols associated with central Mexico also are portrayed on stelae from Maya Tikal. Yet despite these long-distance exchange and diplomatic contacts, there is little evidence that Teotihuacan directly controlled much territory outside the Basin of Mexico. We also have
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
no depiction, symbolic record, or even mortuary context that records an omnipotent ruler at Teotihuacan. In fact, mural art at the site often illustrates a line of well-dressed but similar figures in a procession (Figure 8.35). Little is known about the decline of Teotihuacan in the seventh through tenth centuries A.D. The site was not abandoned, but its size decreased by more than half during this period. Militarism is a prominent theme in art during A.D. 650–750, although perimeter defensive walls were never constructed. Part of the city was burned in the seventh and eighth centuries; the fires appear to have started intermittently. The core of the city—the buildings along the Avenue of the Dead— was burned, as were temples, pyramids, and public buildings throughout the site. Millon has suggested that the conflagration was deliberate and ritu-
Page 355
ally inspired, similar to the earlier desecration of the Olmec stone sculptures. In ancient Mesoamerica, the symbolic destruction of selected monuments or sacred structures was repeatedly associated with the decline and loss of power. After A.D. 750, the enormous pyramids, the avenues and markets, and the memory of the Classic civilization were in decline. Centuries later, the Aztecs referred to the mounds of stone and broken walls as “the place of the gods.” Today, Teotihuacan is one of the most important tourist attractions in Mexico. The magnificent Avenue of the Dead has been reopened and leads past the Pyramid of the Sun to the plaza and Pyramid of the Moon. The pyramids have been restored, and every year thousands climb them in awe of and admiration for the achievements of the ancient Mexicans.
Most [artistic] scenes [at Teotihuacan] show human beings so loaded with clothing and insignia that faces and other body parts are barely visible. Emphasis is on acts rather than actors; on offices rather than office-holders. This, together with the multiplicity of identical scenes, suggests an ethos in which individuals were interchangeable and replaceable. . . . Supreme [Teotihuacan] political authority may not always have been strongly concentrated in a single person or lineage. —George L. Cowgill (1997)
_ Site Teotihuacan
355
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 356
Concept
<
The Mesoamerican Ballgame A ritual game with symbolic and political importance
Figure 8.36 The large I-shaped ball court on the Main Plaza at Monte Albán.
Figure 8.37 A stone relief panel from the ball court at El Tajín.
_ _
tlachtli The Aztec word for their ritual
ballgame.
356
The art historian Paul Kirchhoff drew up a list of cultural traits that generally defined the prehispanic societies of Mesoamerica and distinguished them from other large cultural regions in the New World, such as the Andean area in South America. Among the traits he used to define the boundaries of prehispanic Mesoamerica was a game played with a solid rubber ball. First reported in sixteenth-century Spanish accounts, the ballgame, tlachtli, was a team game played by somewhat different rules from place to place. Ceramic evidence suggests that some form of the ballgame was played in Mesoamerica as early as 1000 B.C., soon after the transition to village life. Ballplayers are depicted in clay figurines from central and western Mexico, as well as sites along the Gulf Coast. The pottery figures are shown wearing such playing gear as knee guards. Archaeological evidence also suggests that the ballgame was played in different ways during its history. For
Chapter 8 Ancient Mesoamerica
example, the Spanish chroniclers witnessed a game played on a court where stone rings were used as goals. Most ball courts from before A.D. 700 do not have such rings of stone. Small, shallow pits enclosed by four earthen retaining walls, known from San Lorenzo, may have served as early playing fields in the Olmec period. The earliest I-shaped ball courts with stone walls were not built until several centuries later (Figure 8.36). This more traditional form, often with sloping playing surfaces, has been found throughout Mesoamerica. (To compare the oval ball court of the U.S. Southwest, see “Snaketown,” Chapter 7, p. 300.) The absence of a ball court at Teotihuacan is a mystery, since ballplaying is depicted on polychrome murals in one of the compounds. Ballplayer figurines also have been found at the site. The Mesoamerican ballgame was not just for the sake of sport. Spanish chroniclers describe the sixteenthcentury version as very rough, played with a ball weighing up to 11 kg (5 lb). In one version, the ball had to be kept in motion and could not be hit with hands or feet. Hips, knees, and elbows were used, and injuries were frequent. In prehispanic times, the game was associated with fertility, death, militarism, and sacrifice. At the site of El Tajín in Veracruz, a stone relief panel in the ball court graphically shows one player stretched over a sacrificial stone, while another is poised with a stone knife ready to be plunged into the chest of the victim (Figure 8.37). Sixteenthcentury accounts also detail the sacrifice of defeated team members. Interestingly, the ball courts at the Maya site of Chichén Itzá and at the Aztec capital of Tenochtitlán were directly adjacent to the tzompantli, the skull rack where the heads of war captives were placed.
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 357
Site
:
Tikal A Maya city in the rain forest of Guatemala Figure 8.38 A map of the core area of Tikal, showing the major buildings and causeways.
ay
ew
s au
Ma u
ds lay
C
North Acropolis
Temple IV Tozze
Great Plaza Temple I
r Cau
seway
Temple II
Temple III
Central Acropolis
M
en
de
zC
au
se
South Acropolis
wa
y
Temple of Inscriptions 0
500 m
Reservoirs
The first adventurers and antiquarians who discovered the overgrown ruins of Maya sites in the rain forests of the Petén assumed them to be the remains of ancient cities with large populations. Yet by the early 1900s, opinion had shifted to the belief that the ancient Maya resided in vacant ceremonial centers, inhabited only by small groups of priests. These individuals were purported to direct rural, peasant populations through periodic rituals at the centers. This shift in perspective, from urban to ceremonial, was the result of the concentration of archaeological investigations at the ceremonial cores of these sites, and of preconceived notions
that tropical forests could not support urban centers. Researchers failed to recognize residential architecture, and they often neglected to look for houses outside the ceremonial cores of the sites. Moreover, since today the Maya practice slash-and-burn cultivation, it was assumed that the ancient Maya used the same technique. Slash and burn is a simple strategy that involves a cyclical process of field clearing, cultivation, and abandonment. Such extensive agriculture was not thought capable of supporting large urban enclaves or high population densities. During the past three to four decades, however, new information has
slash and burn A type of farming in
which the ground is cleared by cutting and burning the vegetation on the spot. The burned vegetation serves as a natural fertilizer. The field is farmed until yields decrease; then it is allowed to lie fallow.
Site Tikal
357
_
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Figure 8.39 In this view of Tikal from the top of Temple IV, only the tallest buildings are visible above the dense rain forests of the Petén: Temple III in the center and Temples I and II to the left.
_ _
Of all the New World states, it is perhaps the Maya that afford us the greatest opportunity for understanding the evolution, operation, and demise of a complex society. No other New World state offers us such a variety of complementary data sets, including eye-witness reports preserved as ethnohistorical documents, hieroglyphic texts that span some 600 years, regional settlement pattern data, linguistic reconstructions, subsistence data, and architectural evolution. The challenge remains for us to integrate all these lines of evidence, to highlight the differences and similarities among them, and to understand the Maya better by learning more about other Mesoamerican states. —Joyce Marcus (1983b)
358
Page 358
revived an interpretation similar to the original one. Two decades of investigations at the site of Tikal (tee-CAHL) by researchers from the University Museum of the University of Pennsylvania included survey and mapping, excavation in selected areas, and the recording of carved stone monuments. The Tikal survey recorded thousands of small mounds in the thick forest vegetation. More common near the center of the site, the mounds decreased in number toward the periphery. A few of the mounds were excavated and found to be residential structures. At Tikal, the quantity of these residential structures demonstrated that the site had held a sizable population. Subsequent mapping of other Maya centers has indicated that the lowland population during the Maya Classic period, A.D. 250–900, was much larger than anyone suspected, considerably larger than it is today. Once researchers recognized the population density of the ancient Maya, new questions were raised concerning agricultural subsistence, craft produc-
Chapter 8 Ancient Mesoamerica
tion, and sociopolitical organization. It is now clear that the ancient farmers of the lowlands used several strategies, such as terracing and ridged fields, for more intensive agricultural production (see “Wetland Fields,” p. 363). Classic Maya communities apparently also had specific occupational groups, such as craftspeople, in addition to the groups of peasants and ruler-priests envisioned in the ceremonial center model. The core of Tikal is situated on a series of low ridges standing roughly 50 m (165 ft) above two swampy areas (bajos) (Figure 8.38). Most of the great structures presently visible, clustered on the rises, date to the later part of the Classic period, the era of most pronounced construction activity at Tikal and in the Maya Lowlands in general. Yet excavations have revealed that many earlier buildings, some contemporaneous with the major Formative period occupation at El Mirador, were encased within the later structures. For most of Tikal’s history, the core of the city was the Great Plaza, al-
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
though it did not dominate Tikal to the extent that the Main Plaza did at Monte Albán. In Tikal, this central open area, laid out before 100 B.C., extends for roughly 1 ha (2.5 acres) and was replastered four times. At the eastern end of the plaza, the Temple of the Great Jaguar (Temple I) rises to 45 m (150 ft), with its crowning roof comb peeking out above the tropical canopy of mahogany, cedar, and chicle (gum) trees (Figures 8.39 and 8.40). A large seated lord was originally painted in an array of colors on the face of the hollow comb. Temple I itself has three high, narrow rooms with carved wooden beams spanning its door frames. The innermost of these beams, or lintels, portrays a Maya lord towered over by a jaguar, which may serve as his protector. Temple I was dedicated in the Late Classic period to Jasaw Chan K’awiil I, the ruler whose sumptuous grave was found beneath its base and whose portrait is found on the lintels and roof comb. Directly across the plaza from Temple I lies the slightly shorter Temple II, probably dedicated to Jasaw’s wife. The north side of the plaza is framed by the North Acropolis, a huge, 100 ⫻ 80 m (325 ⫻ 260 ft) platform that was continually expanded between 200 B.C. and A.D. 550 (Figure 8.41). The University of Pennsylvania team, led by William Coe, intensively excavated this structure, discovering a rich succession of elaborate, elite tombs. Based on their findings, the North Acropolis appears to have been a burial place for a long sequence of Tikal’s rulers. South of the Great Plaza sits Tikal’s greatest palace complex, the Central Acropolis. In total, this complex spreads over more than 1.6 ha (4 acres) and contains a maze of 42 multistory buildings with multiple rooms interspersed with internal courtyards. The elaboration of the room decorations, which included thronelike seats with armrests, implies that Tikal’s lords and their retainers lived and conducted their activities here. Other monumental buildings and important plazas are linked to the Great Plaza by broad, raised causeways called sacbes, or rain forest paths. The Tozzer
Page 359
Causeway (named for a renowned Mayanist) runs west from Temple II to one of the two tallest structures in the pre-Columbian world, Temple IV. Temple IV, some 70 m (230 ft) tall, was built in the middle of the eighth century A.D. by one of Tikal’s last well-known rulers, Jasaw Chan K’awiil’s son and successor, Yik’in Chan K’awiil, possibly to commemorate his father. Tikal has a long history that is marked by two eras of powerful kings and monumental construction (see “Tikal’s Monument Record,” pp. 361– 362). Following the decline of El Mirador at the end of the Formative period, new polities rose to power. Tikal, which was strategically situated to control
Figure 8.40 The Temple of the Great Jaguar (Temple I).
roof comb An architectural feature, frequently carved with glyphs and figures, that is placed on the top of Mesoamerican temples. lintel A horizontal beam of wood or stone that supports the wall above a doorway or window. sacbe The Maya word for a raised
causeway constructed of stone blocks and paved with gravel and plaster.
Site Tikal
359
_
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Figure 8.41 Monumental architectural structures in the North Acropolis at Tikal.
_ _
acropolis (plural acropoli) (Greek) A raised complex of palaces and courtyards, especially in Mesoamerica and Greece.
360
Page 360
major inland trade routes, emerged as the largest and most important Early Classic city in the Maya region. The first ruling dynasty at Tikal, which included several female rulers, exerted power over many of its neighbors, including nearby Uaxactún. Yet there was constant interpolity competition and shifting alliances among the rulers of a network of Maya centers. Tikal’s principal rival, Calakmul (cahlahk-MOOL), forged a series of alliances with nearby centers, including Caracol, and in the mid-500s A.D. defeated Tikal’s ruler. For roughly the next 120 years, Calakmul, which equaled Tikal in size and monumental construction, and its political partners dominated the Maya region. There was little new construction at Tikal during that time, a sign of its fallen fortunes. Jockeying for power, however, did not cease, and eventually, in A.D. 695, the Tikal king Jasaw Chan K’awiil I defeated Calakmul. For the next 100 years, during the reign of Jasaw and
Chapter 8 Ancient Mesoamerica
his successors in the Late Classic, monumental construction reached its peak at Tikal. Under Jasaw and his immediate successors, Tikal’s central Twin Pyramid groups and the most massive acropoli were built, and the site grew to more than 120 sq km (45 sq mi). At its maximum size, bounded by the swampy bajos to the east and west and defensive earthworks to the north and south, the site is thought to have been occupied by 50,000–80,000 people. Around A.D. 800, Tikal began to decline again, little construction was carried out, and few stelae were erected after that date. Whereas a large percentage of Tikal’s excavated house mounds were occupied during the major episode of building, most of the habitation after A.D. 830 was restricted to the site core, and the total site population may have been only 10% of what it was 100 years before. A century or two later, this giant city was largely abandoned by humans and eventually reverted to forest.
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 361
Concept
<
Tikal’s Monument Record The rulers of Tikal
Recent studies in Maya archaeology and epigraphy have demonstrated that Classic Maya society was not egalitarian but, instead, was hierarchically organized. The commemoration of the births, accessions, conquests, and deaths of important rulers was a major theme of their carved stone monuments. One of the most important monuments at Tikal, known as Stela 29, exhibits the earliest known and wellprovenienced inscriptions in the Maya Lowlands with a Long Count date of A.D. 292 (see “Writing and Calendars,” p. 369). For archaeologists, Stela 29 is a rough marker for the beginning of the Maya Classic period, an era during which the Long Count was used on numerous Maya inscriptions (Figure 8.42). The end of the Classic period (A.D. 900) roughly coincides with the cessation of Long Count inscriptions. Stela 29 also includes the earliest emblem glyph (place name), a distinctive sign that identifies either a specific Maya center or the ruling lineage associated with that location. In the Early Classic period, the importance of Tikal and its ruling lineage is indicated by the numerous dated monuments at the site. Although other Maya centers in the immediate vicinity of Tikal also asserted their importance and autonomy by erecting fourthcentury A.D. monuments, Tikal and its neighbor Uaxactún (wah-shak-TOON) raised almost half of these early stelae. Stela 5 at Uaxactún, dating to A.D. 358, includes Tikal’s emblem glyph, an indicator that the former site by that time may have been under Tikal’s jurisdiction. Soon thereafter, several other Petén sites ceased erecting stelae, another signal that Tikal may have expanded its sovereignty early in the Classic period. Precocious political development at Tikal and Uaxactún may have been related to the close ties that these centers
developed with foreign elite. A stela erected in A.D. 378 at Uaxactún portrays a figure in non-Maya attire. Less than 2 years later, Tikal Stela 4 records the accession of a new ruler, Yax Nuun Ayiin I, who is posed facing front rather than in profile, as was typical of earlier Maya rulers. In pose and costume, Yax Nuun resembles Teotihuacan nobles, and the excavation of the tomb thought to have been his included objects suggesting a Teotihuacan connection. This foreign link continued, although perhaps somewhat diminished, during the reign of Yax Nuun’s heir, Siyaj Chan K’awiil II. The latter figure is portrayed on Tikal Stela 31, which commemorates the twentieth year of his rule (Figure 8.43). On the stela, Siyaj Chan appears in Maya regalia, flanked by two subordinate figures wearing outfits more indicative of the highlands. Their shields are adorned with the face of the central Mexican rain deity prominent at Teotihuacan. During and immediately after the reign of Siyaj Chan, Tikal’s influence spread even farther in the lowlands, possibly stretching to distant Yaxchilán (yahchee-LAHN), a Maya center on the Usumacinta River. Tikal’s broad foreign ties and political importance may have stemmed from the site’s location on land that lies between two major drainage systems, which linked the Gulf of Mexico with the Caribbean Sea, interconnecting the Maya Lowlands. Early in the sixth century A.D., a stela at Yaxchilán displayed the site’s own emblem glyph, which may imply that a greater degree of autonomy from Tikal had been secured. Tikal’s declining influence was signaled by a construction slowdown and a long break, from A.D. 534 to 692, in the carving of dated stelae. During the first 59 years of this period, called the hiatus, monument erection practically ceased at many sites in the
Figure 8.42 Stela 29, with a richly dressed noble on one side of the monument and glyphs arranged in a single column on the other.
epigraphy The study of inscriptions.
Concept Tikal’s Monument Record
361
_
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Figure 8.43 Stela 31, showing Siyaj Chan K’awaiil II flanked by two subordinate figures.
Long Count The Classic Maya system of dating that records the total number of days elapsed from an initial date in the distant past (3114 B.C.). The system is based on multiples of 20. emblem glyph A set of Maya hiero-
_ _
glyphs; generally, each emblem glyph is specific to a given Classic Maya city.
362
Page 362
southern Maya Lowlands. Just before the hiatus, a Tikal ruler, Wak Chan K’awiil, was apparently defeated in battle by Caracol (car-ah-COAL), an ally of Calakmul. During the hiatus, even the tombs of the Tikal elite contained meager offerings compared with the exotic and sumptuous items buried earlier. Recent epigraphic breakthroughs stemming from the decipherment of glyphic texts from across the Maya Lowlands support archaeological indications that the power of Tikal’s lords waned during this period. At the same time, another Maya center to the north, Calakmul, Tikal’s principal rival, ascended to much greater influence. But by A.D. 700, political fortunes appear to have shifted again. Late in the seventh century, the most extensive construction episode was begun at Tikal under the ruler Jasaw Chan K’awaiil I, who deliberately harkened back to powerful Siyaj Chan K’awaiil II. Clemency Coggins, of Boston University, has proposed that Jasaw’s accession was timed to occur exactly 256 years after the inauguration of Siyaj. The later Classic period was characterized by a series of politically competitive centers, including Tikal in the central Petén, Calakmul in the north, Palenque in the southwest, Yaxchilán along the
Chapter 8 Ancient Mesoamerica
Usumacinta, and Copán in the southeast. The specific fortunes of these centers, their interconnecting networks of alliance, and the overall degree of political consolidation in the Maya Lowlands appear to have vacillated over time. Yet during this era, stylistic and iconographic conventions were shared across the southern Maya Lowlands. A standardized lunar calendar was employed throughout the region for almost a century (A.D. 672–751). Monuments from various centers indicate that communications between the sites were fostered by elite intermarriage, as well as by military alliances. Archaeological findings suggest that exchange also was an important means of regional integration. Whether the Late Classic period ruling lineage at Tikal ever developed the far-flung influence that Yax Nuun and Siyaj Chan apparently had achieved earlier remains a matter of scholarly debate. Although recent epigraphic findings indicate that the later lords of Tikal did extend their political influence over other Maya centers, those same texts seem to tell of increasing political fragmentation after A.D. 750. By the end of the eighth century A.D., construction at Tikal, and at many other major Maya centers, began to wane. In A.D. 889, the latest monument was erected at Tikal, and the last known Long Count date anywhere was carved in A.D. 909. Tikal’s fate was only part of the significant demographic decline and political breakdown, the so-called Maya collapse, that characterized almost the entire central and southern lowlands during the ninth and tenth centuries.
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 363
Concept
<
Wetland Fields Intensive agriculture in the Maya Lowlands
Fifty years ago, most Mayanists believed that the ancient Maya relied entirely on the slash-and-burn farming practices used in Yucatán today. For example, in 1956, Sylvanus Morley and George Brainerd summarized the prevailing viewpoint as follows: Modern Maya agricultural practices are the same as they were three thousand years ago or more—a simple process of felling the forest, burning the dried trees and bush, planting, and changing the locations of the cornfields every few years. (p. 128) Yet during the past decades, with the completion of archaeological settlement surveys, a large number of sites have been mapped. Often these sites were found to have contained a greater residential population than expected, raising questions about Maya subsistence and population. Most notably, given the long fallow cycles required by slash-and-burn farming, how did the large Classic Maya populations sustain themselves? Dennis Puleston, who surveyed outlying portions of Tikal, provided an alternative hypothesis. Noting the high spatial association between the ruins of Tikal house mounds and the distribution of ramón trees, Puleston suggested that the fruit of the ramón, eaten today as a dietary supplement, may have been a food staple in the past. Although Puleston may have overestimated the ancient dietary significance of ramón, more recent findings have supported his view that the ramón and other fruit trees were tended by the prehispanic Maya and probably contributed to their diet. In 1969, Puleston, collaborating with geographer Alfred Siemens, discovered raised-field complexes during an aerial reconnaissance in the vicinity of a series of Maya sites along the Can-
delaria River in the state of Campeche. Subsequent aerial photographs, onthe-ground checks, and archaeological excavations confirmed that the observed patterns of raised fields and interdigitated canals were constructed before the Spanish conquest. Since the initial work of Siemens and Puleston, these agricultural features, which are similar in function to the prehispanic central Mexican chinampas, also have been noted in other parts of the Maya Lowlands (see “Tenochtitlán,” p. 378). Along the Hondo River in Belize, Puleston’s excavations suggested that the prehispanic raised features were built up by piling floodplain sediments above the natural terrain (Figure 8.44). In other areas, canals were constructed to drain water from swampy and waterlogged zones. In both cases, cultivated fields were established above the high-water table, or floodplain, during the rainy season. Yet, if necessary, these raised or drained fields could receive water supplements from the intervening canals in the dry season. In addition, the periodic cleaning of the canals further raised and naturally fertilized the fields. Although these features could be laborious to construct, the returns were significant, allowing for a much more continuous output than that afforded by slash and burn. The canals within the raised-field systems also could have promoted fish cultivation and facilitated transport and communication. The Maya glyph signifying abundance was the water lily, a key natural floral component of the artificially constructed aquatic raised-field/canal environment. Analyzed pollen samples from the silt of the ancient Hondo River canals have identified maize and cotton as crops grown in the raised-field complexes. Thus subsistence foods as well as nonfood plants were grown.
I’m not particularly interested in ancient objects. This seemingly heretical statement for an archaeologist usually takes aback friends who believe that the best way to entertain me is to show me the local museum. On more than one occasion, I have had to explain that beautiful Classic Maya vases or finely carved jade pendants hold less interest to me—and to many of my colleagues— than the scientific investigation of ideas about why and how ancient cultures like the Maya developed. —Jeremy A. Sabloff (1990)
ramón A tree that grows abundantly
in the tropical forests of the Maya Lowlands and bears an edible fruit, also called breadnut. chinampa (Spanish) An agricultural field created by swamp drainage or landfill operations along the edges of lakes.
Concept Wetland Fields
363
_
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 364
Figure 8.44 An aerial view of a raised-field complex in Pulltrouser Swamp, Belize.
_ _
Indigenous knowledge systems of landscape management need to be studied and evaluated before they disappear forever. In cases where the prehistoric infrastructure has been completely abandoned, investigations using archaeology and agricultural experiments may be able to recover sufficient information on how these systems function. . . . Both contemporary and prehistoric systems may hold the clue to future rural development. —Clark L. Erickson (1992)
364
Significantly, radiocarbon dates from the Hondo River fields place the initial construction of this artificial landscape in the first millennium B.C., before significant population expansion in the Maya Lowlands. Side-looking airborne radar (SLAR) has been used to locate huge areas of suspected raised-field complexes in a large portion of the Maya region. Yet most of these proposed field systems have not been checked on the ground. Several other linear patterns observed by SLAR reflect no more than natural soil formations. Thus the geographic extent of ancient Maya raised-field complexes remains unknown. In several areas, the ancient Maya supplemented extensive slash-andburn farming with the construction of artificial terrace systems. Evidence for these agricultural features, which converted otherwise unfarmable slopes into a patchwork of level plots, has been found both in western Belize and in the vicinity of the Bec River. Although doubts persist about the dating
Chapter 8 Ancient Mesoamerica
and distribution of Classic Maya agricultural systems, there is no longer any question that the Maya employed a great diversity of subsistence techniques, including a range of water management strategies and various kinds of intensive farming procedures.
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 365
Site
:
Palenque A Classic center at the edge of the Maya Lowlands
The Maya and many other native groups in the Americas divided their physical realms into four quarters; each quarter was associated with specific colors, living things, and supernatural forces. Original Maya maps consistently place east at the top, suggesting that east was viewed as the dominant direction. In A.D. 731, a stela was erected at the site of Copán (ko-PAHN), a large center in the southeastern part of the Maya Lowlands. This monument lists the place names of four major Maya centers, each associated with one of the four cardinal directions. Preceding this list is a clause that Joyce Marcus interprets as “four on high,” or “divided into four quarters.” It appears that the Maya associated each section of their world with an important place
0
or center. Not surprisingly, the monument carvers at Copán associated their site with the foremost direction, east. Tikal was associated with the west, and Marcus has suggested that the emblem glyph associated with the south belonged to Calakmul. Palenque (paLEN-kay), a center at the western edge of the lowlands, was linked with the direction north. Although these directions did not correspond closely to the actual geography, the cosmological model, dividing the Maya world between a series of centers, did reflect the political fluidity of the Late Classic Maya realm. This is not to say that the lowland Maya world contained four autonomous centers in A.D. 731. Rather, some authority at Copán perceived these four centers to be of the greatest
Figure 8.45 A map of the principal buildings in the central part of Palenque.
150 m
N
Temple of the Inscriptions
Palace Aqueduct
Ball court
_ Site Palenque
365
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 366
Figure 8.46 An artist’s interpretation of the Temple of the Inscriptions, showing the staircase leading to Lord Pakal’s tomb below.
sarcophagus A stone coffin, usually
_ _
decorated with sculpture and/or inscriptions.
366
importance. The listing of multiple centers speaks to the political fragmentation of the Classic period Maya world and the frequent rises and falls in political fortune of the Maya centers that were part of that landscape. Palenque is situated on a series of rolling hills overlooking the vast Gulf Coast plain approximately 50 km (30 mi) from the Usumacinta River, in the state of Chiapas, Mexico. The site was occupied early in the Classic period but remained relatively small until the seventh century A.D. Following a military defeat of Palenque’s ruler by Calakmul in A.D. 611, a major building boom was started by a powerful lord who, inscriptions indicate, came to power in A.D. 615 and may have been named K’inich Janaab’ Pakal I (“shield”). During Pakal’s rule, Palenque also grew in size and expanded its authority over the neighboring region. Palenque emerged as a major power, and the site’s emblem
Chapter 8 Ancient Mesoamerica
glyph appeared. The major structures of the site today were begun after Pakal’s accession (Figure 8.45). Lord Pakal was buried in an elaborate tomb beneath the pyramid that supports the Temple of the Inscriptions (Figure 8.46). The tomb was discovered by Alberto Ruz Lhuiller (in 1949–1958). Several years later, after clearing a staircase that led deep under the pyramid, Ruz found a large stuccoed crypt. The huge stone coffin, or sarcophagus, was topped with a superbly carved lid depicting the dead ruler falling into the underworld (Figure 8.47). Beneath the lid were the remains of Pakal, wearing beads and a mosaic mask of jade, surrounded by other treasures (Figure 8.48). Epigraphic studies of the stone lid of the sarcophagus and other inscriptions at Palenque have led to the reconstruction of the long dynastic history of the site. Pakal was succeeded directly
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
by his son K’inich Kan Balam II, who in a relatively short reign continued the massive construction at the core of the site. Following Kan Balam’s death early in the eighth century A.D., his brother Kan Joy Chitam II replaced him. During his rule, Palenque’s realm reached its greatest extent. Kan Joy also expanded Palenque’s so-called Palace structure by erecting an unusual four-story tower in veneration of his father (Figure 8.49). During winter solstices, the sun seems to set directly behind the Temple of the Inscriptions when viewed from the top of the Palace tower. After the death of Kan Joy, the site and its inscriptions began to wane. A series of three rulers with brief and possibly broken reigns followed. The last recorded accession date of a ruler in Pakal’s lineage was A.D. 764. The last Long Count date known from Palenque (corresponding to A.D. 799) occurs on an inscribed pottery vessel and records the accession of a ruler with a central Mexican–style name (non-Maya), suggesting that the site may have been taken over or influenced by foreigners late in the eighth century A.D. This ended a long recorded sequence of rule that included at least one female ruler. Palenque was one of the earliest Maya centers to experience collapse. Monumental construction and the textual record waned during the latter half of the eighth century and ceased early in the ninth century. Pomona, a small former dependency of Palenque, displayed its own emblem glyph in A.D. 771, suggesting that it had achieved independence. Palenque was also one of the first sites to show increasing ties to Gulf Coast or central Mexican elements. In parts of the Classic Maya world, this external bond often was associated with the cessation of monumental building and the erection of stelae. On a stela that was raised at the site of Seibal (sayBALL) in A.D. 889, Palenque and Copán were no longer considered among the four primary centers of the Maya region. By then the realm of the Classic Maya had shrunk to little more than its former core around Tikal. Although Palenque at its height was clearly part of the Classic Maya
Page 367
Figure 8.47 Inscriptions on the lid of Lord Pakal’s coffin.
_ Site Palenque
367
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 368
Figure 8.48 The exquisite jade mask that was found in the sarcophagus, next to Lord Pakal’s left shoulder.
Figure 8.49 The Palace at Palenque. The multiroom palace sits at the core of Palenque’s complex of civic-ceremonial structures. The most distinguishing feature of the palace is a four-story tower, which provides a superb view of the city and the surrounding countryside.
_ _
368
Chapter 8 Ancient Mesoamerica
cultural realm, the site’s distinctive layout and architecture reflected its location at the western periphery of the lowlands. Compared with the imposing appearance and vertical thrust of the architecture at Tikal and other Classic Maya centers, Palenque’s major structures were broader and more dispersed, giving the site’s civic-ceremonial center a more horizontal flow. Palenque also lacked a corpus of freestanding sculptured monuments (stelae and altars). Instead, hieroglyphic inscriptions generally were written on stone panels or plaster and incorporated directly into buildings.
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 369
Concept
<
Writing and Calendars
Introductory glyph
“Jaguar, Lord of Yaxchilán”
Distance number: 5 days, 16 uinals, 1 tun, 15 katuns (= 108,685 days) Calendar Round position
The earliest Mesoamerican writing appeared more than 2500 years ago, before the rise of the state or the existence of urban centers (see “Carved Stones and Early Writing,” p. 341). Four different prehispanic Mesoamerican writing systems are known: the Maya, the Zapotec, the Mixtec (MEESH-tec), and the Aztec. Each appears to include a mixture of pictographic, ideographic, and phonetic elements. The Mixtec and the Aztec are known from late prehispanic texts written on prepared bark paper or deer skin; the older Zapotec system is preserved on inscriptions carved in stone. The Maya writing system has been preserved in late prehispanic folding books, as well as in a large body of texts on stone, pottery, and wall paintings (Figure 8.50). The Maya writing system is the best known of the four, and major breakthroughs in decipherment have come during the past two decades. At first, the Classic (A.D. 250–900) Maya inscriptions were thought to relate exclusively to astronomy and calendrics, because these subjects constituted the first portions of the texts to be deciphered. Today, epigraphers recognize that the principal theme of Maya inscriptions is a political or dynastic one. As discussed earlier (see “Tikal’s Monument Record,” p. 361), scholars can now chart the military claims and political alliances of specific Maya rulers and compare these textual data with the archaeological and architectural record. Carved stones have been deciphered, which commemorate significant events—birth, autosacrifice, accession to rule, marriage, military victory, death—in the lives of the Maya lords and their kin. Often the text helps frame a stylized scene of the commemorated event. The blocky hieroglyphs are presented in double columns intended to be read
Long Count date
The Maya numerical system and Calendar Round
Calendar Round position
“Bird Jaguar”
from left to right and top to bottom. As with ancient Egyptian monuments, the messages are generally brief and composed largely of nouns and verbs. Prehispanic texts are closely linked to the pre-Columbian calendar. Classic Maya inscriptions are distinguished by their frequent use of the Long Count calendar, a system capable of tracking extended cycles of time (Figure 8.51). Although the Maya did not invent the Long Count, they refined and used it to the greatest degree. The smallest Long Count unit is the day, or kin. The second Long Count unit, composed of 20 kins, is the uinal. The third unit, the tun, consists of 18 uinals or 360 kins.
Figure 8.50 An example of Maya hieroglyphs.
pictograph A written or painted symbol that more or less portrays the represented object. ideograph A written symbol that
represents an abstract idea rather than the sound of a word or the pictorial symbol of an object (pictograph). phonetic Pertaining to the sounds of
speech.
Concept Writing and Calendars
_
369
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 370
Positional values
0
1
4
5
11
19
20
126
1092
36,102 1,368,080
x
144,000
x
7200
x
360
x
20
x
1
baktun katun tun uinal kin
days
Kn ife Mot ion
nt Fli
Figure 8.52 The 260-day Sacred Almanac for the Aztecs. The 20 named days intermesh with the numbers 1–13.
_ _
y
Why do you think the Maya stopped using the Long Count calendar?
370
t elo Oc le Eag
Alligator er Flow in Ra
r ate W g o D Monkey Gra ss Re ed
Vulture
ent ard Liz
Serp Ho us e Win d
Death's head Dee r Ra bb it
Figure 8.51 The Maya numbering system and Long Count units. The base-20 system used only three symbols, the stylized shell for 0, a dot for 1, and a bar for 5. The numbers were arranged vertically (rather than horizontally, as in our Arabic numeration), with the lowest values at the bottom.
Above this third cycle are the katun and the baktun, consisting of 20 and 400 tuns, respectively. In accord with all prehispanic Mesoamerican numeration systems, the Long Count had a vigesimal structure, one based on multiples of 20. The only adjustment necessary to approximate the solar year was a change in the third unit, or tun, from 20 to 18 uinals. The Long Count system enables scholars to date precisely many of the Classic Maya inscriptions. Long Count dates were calculated by the Classic Maya and their Mesoamerican forebears, who established the system by adding the total amount of elapsed time to a base date from which they began their calendar. Until this starting point was calculated, Long Count dates could not be equated with our calendric system. Some scholarly disagreement still exists, but most re-
Chapter 8 Ancient Mesoamerica
searchers place the base date for the Long Count near the end of the fourth millennium B.C. (with 3114 B.C. the most widely accepted date). In addition to the Long Count, the Maya and other Mesoamerican peoples used the Calendar Round. The main element of this calendar was the Sacred Almanac, or 260-day count (Figure 8.52). Primarily religious and divinatory in function, the almanac can be visualized as having been composed of two integrated “cogged wheels,” or cycles of numbers; one cycle of numbers ran from 1 to 13, and the other cycle consisted of 20 named days. In general, these named days were similar across Mesoamerica; however, some differences did occur from group to group. The combination of a number and a day name formed a unit that would not recur until 260 days had elapsed, at which point the cycle began again. The numbers were usually expressed in typical Mesoamerican bar-and-dot fashion, with a dot equivalent to 1 and a bar equivalent to 5. A second element of the Calendar Round was the 365-day Vague Year. This year was divided into 18 “months” of 20 days each. An extra period of 5 days was added at the end of the year. Although the Classic Maya were aware that the tropical year actually was 365.25 days, they did not use a leap-year correction. The permutation of the Sacred Almanac and the Vague Year produced an 18,980-day Calendar Round that cycled back to its original start every 52 solar years. The Maya believed that significant moments in time had their links with divinities and astrologically based auguries, and these associations were known and recorded. The Maya viewed time in a cyclical fashion. Points in time and the events assigned to them were not thought to be unique; the past, present, and future often could be woven together in prophecy and divination. Given the intricacy of Classic Maya writing and calendrics and their importance in the politico-religious sphere, it appears that this special knowledge was generally held by only a relatively small segment of the population.
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 371
Site
:
Tula Capital city of the Toltec
When the fifteenth-century Aztec ruler Itzcoatl purged the state archives, burning older documentary records, he greatly limited our ability to reconstruct the late prehispanic history of central Mexico. Fortunately, some lists of rulers, called king lists, and other fragmentary accounts of the previous periods survived. In conjunction with archaeological discoveries, these records enable us to begin to understand events in the central highlands between the decline of Teotihuacan, beginning around A.D. 700–800, and the establishment of the Aztec city of Tenochtitlán in A.D. 1325. Sixteenth-century Spanish clerics and men of letters, curious about prehispanic civilization, inquired about the native history of Mexico. Repeatedly, their Aztec informants attributed the origins of their society to a semilegendary race known as the Toltecs (TOLL-tecs). According to these parthistorical, part-mythical accounts, the Toltecs were great warriors but also peace-loving people. These mythic histories also describe the Toltecs as innovators, builders, and craftspeople, who cultivated giant ears of corn and colored cotton that did not have to be dyed. Their capital was supposedly an architectural masterpiece known as Tollan (toll-LAHN). The Aztecs traced their right to rule directly to their Toltec ancestry. Like many such accounts, the origin myths of the Aztecs blended historical reality with fiction and distortion. Some of the stories did contain elements of truth, which enabled Mexican ethnohistorian Wigberto Jiménez Moreno to persuasively suggest that the Toltec capital was an actual place near the modern town of Tula de Allende (TOO-lah de ah-YEN-day), in the state of Hidalgo. Over the past decades, several archaeological pro-
jects at Tula have substantiated Jiménez Moreno’s identification. Tula is situated at the northern fringe of Mesoamerica, about 65 km (40 mi) northwest of modern Mexico City. The site had little significant occupation when Teotihuacan was partially burned and abandoned. The collapse of that great city, nearly coincident with the demise of monumental building at Monte Albán, marked a major transition in the highlands of
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the Toltecs, see the Internet exercises on your online learning center.
Ball court
Ball court
Central altar
Principal pyramid N
0
Figure 8.53
100 m
_
A map of the major buildings in the central part of Tula. Site Tula
371
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 372
Figure 8.54 An artist’s reconstruction of a residential compound at Tula, showing covered living spaces and small open courtyards, or patios.
After the fall of Teotihuacan, various warlike elites, flourishing different symbols of political legitimacy, appropriated the city’s patrimony. These successor states contested one another’s spoils and spread out in various directions in search of new horizons. For a brief time, the center of gravity within the Mesoamerican heartland shifted northward, apparently to Tula in Hidalgo, outside the Valley of Mexico. Tula became the capital of a Toltec domain, less an integrated "empire" than an epicenter of groups of warriors, traders, cultivators, and priests employing the Toltec name and symbols as charters for conquest and colonization. —Eric R. Wolf (1982)
Atlantean column A carved human
_ _
figure serving as a decorative or supporting column.
372
Mexico. Following the fall of Teotihuacan, no single dominant community emerged in the Central Highlands to control the Basin of Mexico. Tula rose rapidly in this politically fragmented landscape. Soon after A.D. 800, a substantial occupation was present at the site (Figure 8.53). Even at its height, Tula was considerably smaller than Teotihuacan or the later Aztec city of Tenochtitlán. Tula also lacks any evidence of urban planning outside its civic-ceremonial core. It reached its maximum size and its greatest influence between A.D. 950 and 1150: 13–16 sq km (5– 6 sq mi) and about 40,000–60,000 people. The residential compounds at Tula are not as large as most of those at Teotihuacan, although they do appear to house more than a single nuclear family (Figure 8.54). The major structures at the site, including two large ball courts, were built during this period. Although the monumental architecture at Tula was far less grandiose than at Teotihuacan, a series of squat but substantial temple pyramids were erected. The most impressive and best-known remains at Tula are the sculpted Atlantean columns, which stand 4.6 m (15 ft) high (Figure 8.55). The military theme depicted by these giant basalt figures is also evident in a series of carved stone relief panels showing jaguars, eagles eating
Chapter 8 Ancient Mesoamerica
hearts, and coyotes, all creatures perhaps associated with specific warrior groups in the city. Trade, the procurement of raw materials, and craft production were very important economic activities at Tula, as they were at Teotihuacan and Tenochtitlán. Containers of travertine, a white sedimentary stone, were made at the site, and sixteenth-century accounts describe many other fine, highstatus goods produced by the Toltec craftspeople. As at Teotihuacan, areas of heavy obsidian use and/or production have been recorded at Tula (Figure 8.56). Most of the obsidian was derived from the same source that furnished workshops at Teotihuacan. Tula borders on arid regions to the north, where little can be grown except maguey. This plant is an important source of fiber, needles, sap, and other products. In ancient Mesoamerica, maguey fibers were raw material for cloth, clothing, nets, and bags. The abundance of spindle whorls at Tula indicates that the spinning of these fibers into twine also was an important economic activity. The presence of pottery and marine shells from the Pacific Coast further documents the extent of exchange with peoples in northern and western Mexico. Both historical and archaeological sources agree that Tula had lost much of its influence before A.D. 1200. Exca-
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 373
looting in earlier monumental structures and even in the residential compounds. Colonial-era texts of the Spanish friar Bernardino de Sahagún allude to this looting, citing the removal of Toltec artifacts from the ground. Various written accounts provide a more intriguing depiction of Tula’s end, but their accuracy is unconfirmed. Although the accounts themselves are somewhat contradictory, the story ties the fall of Tula to a conflict between cosmic forces. The Mesoamerican cosmos was frequently symbolized as a duality. Quetzalcoatl, the representative of day, light, traditional religion, and good, was opposed to Tezcatlipoca, the embodiment of night, darkness, chicanery, and evil. In these accounts, Quetzalcoatl (“the feathered serpent”), or a ruler associated with him, is tricked, disgraced, and forced to leave the city, an event resulting in Tula’s downfall. Both archaeological and historical sources indicate that following the collapse, the Basin of Mexico and its northern fringes once again fragmented into a series of small polities, each controlling only its immediate vicinity. It was during this era of political decentralization that the Aztecs began their ascent to power.
0
5 cm
Figure 8.56 Two obsidian blades and an exhausted core from Tula.
Figure 8.55 One of the large carved Atlantean columns at Tula.
vations at Tula by Mexican archaeologist Jorge Acosta revealed evidence of burning and disruption at all the major structures; however, the specific date of this destruction was not determined. A recent project by Richard Diehl found that Tula was at least partially abandoned by A.D. 1200, although the extent of the collapse is unknown. Diehl’s study also recorded a sizable later Aztec occupation at Tula. Acosta found extensive evidence for late prehispanic _ Site Tula
373
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 374
Site
:
Chichén Itzá The most magnificent late Maya center in Yucatán
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on Chichén Itzá, see the Internet exercises on your online learning center.
The collapse of major centers in the central and southern Maya Lowlands during the ninth and tenth centuries A.D. coincided with an episode of development and population increase in the drier northern fringe of the Yucatán Peninsula. Several clusters of larger centers were established in the northern lowlands, each marked by its own distinct architectural style. The most widespread of these, the Puuc (pook) style, is identifiable by a
Figure 8.57 Puuc-style architecture at Chichén Itzá. This architectural style was centered in the only hilly region of the northern lowlands. (Puuc is the Maya word for hills.)
cenote The Maya word for a sinkhole,
_ _
a natural well in the Yucatán that provides water for drinking and bathing.
374
Chapter 8 Ancient Mesoamerica
mosaic of limestone masonry covering a rubble core (Figure 8.57). The earliest construction at the site of Chichén Itzá (chee-CHEN eet-ZAH), in the ninth century, is characterized by this style. Chichén Itzá, one of the easternmost of the Puuc sites, was strategically situated toward the center of the Yucatán Peninsula, near two large natural sinkhole wells, or cenotes (Figure 8.58). The limestone bedrock of northern Yucatán is very porous, and rainfall soaks
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 375
Figure 8.58 The sacred Cenote of Sacrifice at Chichén Itzá.
Figure 8.59 A map of the central part of Chichén Itzá, showing the principal buildings and causeways.
Sacred W
ay
Cenote of Sacrifice
Ca
use
wa
Ball
cou
rt
y
Tzompantli Temple of the Warriors Ball court Castillo
Ball court
The Ma
Ball court
use
wa
y
Ball court
rket
Ca
immediately into the ground. There are no rivers and very little surface water in this area. The cenotes in the limestone landscape are a very important source of drinking water. The flowering of the Puuc centers is associated with ceramic and architectural elements from the non-Maya areas of Mexico, perhaps the Gulf Coast. The archaeological indications of these exotic connections fit well with later written accounts suggesting the role of “Mexicanized” Maya traders in the rise of the northern Yucatán centers during the Late Classic period. Their influence was related to a shift in long-distance trade routes and involved an expansion of circum-Yucatán sea transport at the expense of the old Petén River route across the Petén lowlands. Late in the tenth century A.D., a second, more opulent construction boom led to the rebuilding of the center of Chichén Itzá in an architectural style more similar to that found at Toltec Tula (Figure 8.59). During this later construction phase, Chichén Itzá grew to its maximum size, more than 5 sq km (2 sq mi). The new architecture included colonnades, extensive I-shaped ball courts (the largest in Mesoamerica) (Figure 8.60), relief carvings of prowling jaguars, carved warrior figures, carved tzompantlis (skull racks; Figure 8.61), and fascinating, sculpted chac mools (Figure 8.62). The hands on these carved human
N
0
100 m
_ Site Chichén Itzá
375
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Figure 8.60 The large ball court at Chichén Itzá. The Mesoamerican ballgame has a long history; however, the specific form and size of the courts varied across time and space.
_ _
Page 376
figures surround a bowl-shaped impression at the navel, where sacrifices or offerings were placed. Some archaeologists have suggested an actual central Mexican Toltec presence at Chichén Itzá on the basis of ancient legends that tell of the god Quetzalcoatl from Tula arriving in Yucatán in A.D. 987. This interpretation is supported, perhaps, by murals in a temple at Chichén Itzá depicting war-
riors (assumed to be Toltecs) arriving in canoes and doing battle with the local populace (Figure 8.63). Yet the same textual sources that tell of foreign invaders in Postclassic Yucatán note that they spoke a Mayan language, suggesting that they were not actually central Mexican Toltecs. Although the nature of the relationship between these two centers is still unresolved, recent interpretations favor a
Figure 8.61 The tzompantli (carved skull rack) at Chichén Itzá depicts human skulls impaled on poles.
376
Chapter 8 Ancient Mesoamerica
pri31057_ch08.pp320-389.qxd
Figure 8.62 Itzá.
10/1/09
8:12 AM
Page 377
A chac mool at Chichén
more dynamic and interactive interconnection that saw goods and information flow in both directions. The later construction at Chichén Itzá reflects a strong Maya influence, with veneer masonry, mosaic facades, and the use of vaults, and buildings are generally more finely made and larger than similar buildings at Tula. The focus of the new construction was a huge 12-ha (30-acre) plaza area. For a century, Chichén Itzá was the primary locus of truly monumental construction in northern Yucatán. The scale of building at the site, including 15 large plaza groups and numerous ball courts, is not found elsewhere. During that era, Chichén Itzá was the largest, most powerful Maya center in the region. Major construction at the site slowed during the eleventh century, and by A.D. 1100, Chichén Itzá had lost its power and prestige. Some buildings were abandoned, and sculptures were destroyed. Later documents describe Chichén’s end by conquest. As with the Classic Maya sites to the south, it is not clear if violence caused the downfall of Chichén Itzá, or if the destruction took place only after the rulers had lost their power and authority. During Chichén Itzá’s ascendancy, and even after its demise, the cenotes remained important sites of sacred offerings. A sacbe (stone causeway) ran directly from the site’s main plaza to one of the cenotes. Offerings of all kinds from across the Maya region, including Classic period items that may have been looted, were cast into this sa-
cred natural well, appropriately named the Cenote of Sacrifice. According to Bishop Diego de Landa, a Spanish cleric who chronicled life in sixteenthcentury Yucatán, men were thrown alive into the well as a sacrifice to the gods during times of drought. Indeed, the dredging of the sacred cenote turned up human bone, as well as prehispanic copper, gold, and jade. According to sixteenth-century accounts, items to be sacrificed were first painted blue. This textual description has been supported by the recent analysis of a small preserved mass of copal incense wedged in a small pot that had been dredged from the cenote. Dean Arnold, of Wheaton College, and collaborators at the Field Museum in Chicago documented the presence of Maya Blue and one of its components (the clay palygorskite) in the preserved wedge of copal. Maya Blue apparently was made on the side of the cenote by heating its components over low fires sustained by copal. The blue pigment was then applied to objects and people before they were dispatched into the sinkhole. At the base of the cenote, a 4.5–5.0-m-(14–15 ft)-thick layer of blue silt was found right above the bottom. This layer of blue silt undoubtedly was formed as Maya Blue pigment washed off the sacrificial offerings after they were tossed into the cenote.
Figure 8.63 Warriors arrive in canoes in this mural from the Temple of Warriors.
chac mool (Maya) A life-size stone figure in a reclining position, with flexed legs and head raised and turned to one side. vault An arched structure of masonry
that forms a ceiling or roof. The construction of vaults with corbelled, or stepped, ceilings was a common building technique of the Maya. Maya Blue A steadfast blue pigment made by fusing an extract from the plant indigo with a fine white clay, palygorskite. The Maya and other Mesoamerican peoples applied this pigment on a range of materials, including ceramics, sculptures, and murals.
_ Site Chichén Itzá
377
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 378
Site
:
Tenochtitlán The capital city of the Aztecs: the Venice of the New World
Figure 8.64 Recent excavations at Tenochtitlán’s Templo Mayor. Today, this Aztec site lies at the center of downtown Mexico City. After the Spanish conquest, Mexico City rapidly grew up over the ancient Aztec center, and materials from the upper level of the Templo Mayor were used in the construction of Mexico City’s main cathedral. w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the Spanish defeat of the Aztec ruler Montezuma, see the Internet exercises on your online learning center.
Chichimec A term loosely applied
to the peoples who lived beyond the northern limits of Mesoamerica; nomadic people, considered to be uncivilized barbarians. petty state A small, socially stratified
_ _
political unit prevalent in Mesoamerica at the time of the Spanish conquest.
378
The Aztec civilization has aroused scholarly interest since the discovery and conquest of Mexico by the Spanish in the early sixteenth century. Although written accounts of Aztec society are rich and full of descriptive detail, the archaeological record is somewhat impoverished, in part because the Aztec capital of Tenochtitlán (ten-noch-teet-LAN) is buried under the bustling metropolis of Mexico City. In February 1978, electrical workers were digging trenches for cables in Mexico City when they found a large stone sculpture. Located 2 m (6.5 ft) below the surface, this sculpture depicted the Aztec female lunar deity Coyolxauhqui (koh-yohl-SHAW-kee), who according to myth was slain and dismembered by her very important brother Huitzilopochtli (wee-tseel-oh-POHCH-tlee), the war god. This discovery revived na-
Chapter 8 Ancient Mesoamerica
tional interest in the Great Temple of the Aztecs, the Templo Mayor: a twin temple dedicated to Huitzilopochtli and Tlaloc, the rain deity. The accidental discovery initiated a long-term research project that has added greatly to our knowledge of this city and the temple that was the physical and psychic heart of the Aztec world (Figure 8.64). The Aztecs, who referred to themselves as Mexica (may-SHE-kah), had humble origins. The Mexica were Chichimecs (CHEE-chee-mex) from the northern desert who did not arrive in the Basin of Mexico until the middle of the thirteenth century A.D. Although the specific cause of this population movement is not known, it appears to have coincided with an episode of drought in the north, which may have made that already dry area particularly unsuitable for agriculture. Alternatively, the Mexica and other northern groups may have been attracted to the economic and military opportunities afforded by the rise of Tula on Mesoamerica’s northern frontier. When the Mexica arrived, more than 50 small, autonomous petty states were present in the Basin of Mexico. At first, the Mexica had little choice but to settle on lands controlled by others. The established basin populations viewed the Mexica as barbarian heathens; yet they also recognized their military prowess, important during a period characterized by military conflicts and competition. The Mexica’s migrations did not end after their arrival in the basin. Feared and despised by more powerful and urban groups, they were repeatedly driven away. They took asylum on lands owned by Culhuacan, a relatively powerful petty state with ancestral ties to the ruling families at Tula. According to legend, the Mexica served the Culhua well as mercenaries. Even-
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
tually, they asked for the hand of the daughter of the ruler of Culhuacan to strengthen their bloodlines through marriage. Unfortunately, one of their gods, Huitzilopochtli, spoke to the Mexica priests and asked them to sacrifice the girl. The Culhua ruler, invited to a ceremony to dedicate his daughter as a goddess, instead found a Mexica priest dressed in the flayed skin of his offspring. His wrath again forced the Mexica to retreat. At the start of their journey, the Mexica had been told by the same god that they would find an eagle perched atop a cactus plant to mark the end of their wanderings. Fleeing into the swamps of Lake Texcoco, one of the principal lakes of the basin, they found the eagle and cactus (a scene now immortalized on the flag of Mexico) on a nearby small, marshy island. The Mexica settled here and erected a temple for Huitzilopochtli. They named the settlement Tenochtitlán (“place of the fruit of the prickly pear cactus”). When Tenochtitlán was founded in A.D. 1325, the island lacked most construction materials (such as wood and stone), suffered periodic flooding, and was abundant only in insects. But the location offered several advantages. Water fowl and fish were plentiful, and the freshwater lakeshore was amenable to intensive chinampa agriculture (Figure 8.65). These farm plots, created by draining standing water and raising the surface of the swamp with soil and vegetation mats, became a major source of food production for the Aztecs. The little island was also in a prime location for controlling transportation in the basin. In ancient Mesoamerica, transport was exclusively by foot or canoe; there were neither beasts of burden nor wheeled vehicles (although toys with wheels were made). Water transport saved time and energy, especially when heavy loads were involved. Canoes plied the waters of the lake as the most efficient means of moving goods and materials between communities. In the less than two centuries between its founding and the Spanish conquest, Tenochtitlán developed from
Page 379
a small island town to the largest and most powerful city in all of Mesoamerica. During its first century of existence (A.D. 1325–1428), the settlement was subordinate to more important basin centers. During this phase, the Aztec constructed chinampas, encouraged immigration to Tenochtitlán, and continued to gain renown for their mercenary skills. Their rulers strategically intermarried and allied themselves with more established elite in the region. By 1428, Tenochtitlán was powerful enough to join with the petty states
Figure 8.65 Part of an earlysixteenth-century map of chinampas in Tenochtitlán. The chinampas are separated by water, and the houses of the owners, whose names are given in Spanish and Aztec hieroglyphs, are situated on squares of solid ground. Major canals are represented by wavy lines, and footprints indicate streets.
_ Site Tenochtitlán
379
10/1/09
8:12 AM
Figure 8.66 The Triple Alliance controlled large parts of western Mesoamerica, although some groups were never defeated and maintained their independence.
Page 380
HUAST ECS
Tributary province Strategic province Provincial capital
CH
pri31057_ch08.pp320-389.qxd
CS ME HI IC
Gulf of Mexico
ME TZ TIT N LA
Enemy states
Tenochtitlán TLAXCALANS TARASCANS
YOPES
PACIFIC OCEAN
N Lake Zumpango Lake Xaltocán
Lake Texcoco
Texcoco
Tlatelolco
Tlacopan
Tenochtitlán
Lake Xochimilco Volcano Capitals of Triple Alliance Other towns Causeway Nezahualcoyotl’s Dike
_ _
0
380
TECS
100 km
0
Figure 8.67 The Basin of Mexico during Aztec times. Settled on an island in the southern part of the lake, Tenochtitlán was connected to the mainland by a series of causeways.
M IX
Chapter 8 Ancient Mesoamerica
Lake Chalco Iztaccihuatl
16 km
Popocatepetl
of Texcoco and Tlacopan (tlah-coPAHN), forming a Triple Alliance that soon vaulted the Aztec capital to a supreme political position in the Basin of Mexico (Figure 8.66). In the years before the Spanish conquest, Tenochtitlán consolidated this political dominion while its armies defeated and demanded tribute from foreign polities across western Mesoamerica and as far away as Guatemala. In 1473, Tenochtitlán defeated its neighbor city of Tlatelolco (tlah-telOL-co) and incorporated the latter, including its giant marketplace, into the metropolis. The dual city was linked to the mainland by three large causeways that ran from the lakeshore to the central ceremonial district of the capital (Figure 8.67). When the Spanish arrived, Tenochtitlán was a city of roughly 150,000–200,000 people. The Spanish leader Hernán Cortés was greatly impressed by the size and grandeur of Tenochtitlán, calling the Aztec city “another Venice.” Early Colonial period maps and accounts indicate that the basic residential unit at Tenochtitlán was smaller
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 381
Figure 8.68 A page from an ancient Mesoamerican codex. Aztec rulers of Mexico recorded the tribute they exacted from conquered peoples in books called codices. This page from the Codex Mendoza shows the name symbols of tributary towns and the kinds and amounts of tribute they owed.
_ Site Tenochtitlán
381
10/1/09
8:12 AM
Page 382
N
pri31057_ch08.pp320-389.qxd
Shrines of Tlaloc and Huitzilopochtli
Temple of Quetzalcoatl
Skull rack
Ball court
0
Figure 8.69 The plan of the ceremonial core of Tenochtitlán. The shrines to Tlaloc and Huitzilpochtli sit atop the Templo Mayor.
_ _
pochteca A privileged, hereditary guild of long-distance Aztec traders.
382
200 m
than the apartment compounds at Teotihuacan. Most household compounds at Tenochtitlán were occupied by nuclear or joint families, each with its own direct access to the streets and canals of the city. Individual compounds were more easily differentiated, permitting public displays of status and wealth. In Aztec society, advances in social position could be achieved through military or economic successes. Numerous craftspeople (including featherworkers, lapidaries, and reed-mat makers) were encouraged to settle in the city, adding to its importance as a commercial center. At that time, many women who were part of Aztec society engaged in some aspect of textile manufacture. The high volume of some finished goods in Tenochtitlán may have served to undermine local craft production at some outlying communities in the basin. Tenochtitlán also obtained large quantities of exotic
Chapter 8 Ancient Mesoamerica
resources and finished goods as tribute payments from defeated polities (Figure 8.68). In general, nearby polities gave foodstuffs, wood, woven mats, and heavier, more utilitarian items; more distant states sent rarer, more valuable goods. Captive states were typically instructed to send local materials. Other exotic goods were brought back to the city by the pochteca, an association of long-distance traders. The pochteca also returned to Tenochtitlán with valuable military or political information. The Aztec tribute demands were heavy, and noncompliance was dealt with harshly. It is not surprising that Cortés’ small band of adventurers received invaluable military and logistic support from Indian groups that despised the Aztecs. Obviously, Cortés and his few hundred men would not have defeated the Aztec armies as quickly and easily without this native assistance. Once victorious, the Spanish forced the conquered Indians to raze the ceremonial core of Tenochtitlán (Figure 8.69), replacing the temples and shrines with churches, government buildings, and palatial residences. Mexico City rapidly grew up over the ancient Aztec center. The upper level of the Templo Mayor was dismantled, and the materials were used to build Mexico City’s Main Cathedral. During the Colonial era, Spanish power in Mesoamerica was strengthened by occupation of the same place as the prior Aztec civic and sacred authority. Eventually, in response to several destructive floods, the once productive lakes were largely drained by the Spanish. The Colonial authorities could not adequately manipulate the complex networks of dikes and canals that had been built and used by the Aztecs. The dry lake bottom now serves as the soft foundation for Mexico’s earthquake-prone capital.
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 383
Concept
<
Aztec Markets Cacao, currency, and “profit” in ancient Mexico Figure 8.70 Aztec market scenes from the Codex Florentino. Top left: a produce market with neatly arranged wares; top right: King Ahuitzotl receiving shells, jaguar skins, plumage, jade, and cacao from the coast; bottom left: members of a slave family wearing bars across their necks as a sign of bondage; bottom right: a merchant from the coast bartering for cloth, gold ornaments, copper, obsidian tools, and maguey fiber rope from the highlands.
In today’s market economies, people try to minimize costs and maximize returns to make a profit. Anthropologists distinguish this type of market endeavor from reciprocal and redistributive exchange networks (see “The Archaeology of Exchange,” Chapter 7, p. 281). Reciprocity refers to face-toface exchanges between known participants, such as relatives or trading partners, in which the specific return is not predefined (gift giving, for example). Redistribution is the accumulation and later dispersal of goods from a central place or by a central person (the income tax system, for example). Aztec marketing, with its network of marketplaces, various currencies, and numerous participants, clearly fits within the conventional definition of a market system (Figure 8.70). The central market at Tlatelolco, Tenochtitlán’s sister city, was the largest market in a network of centers that were situated across the Basin of Mexico. The Tlatelolco market alone served 20,000–25,000 people on a normal day,
and twice as many every fifth day, when it was the focus of a cycle of changing regional markets. Many exchanges in the market were made by bartering. Others were facilitated by accepted currencies. Cacao beans, used to make a favorite drink of the nobility, were the most common form of money, but large cotton cloaks, quills filled with gold dust, stone and shell beads, and copper beads and axes also were employed. Cacao beans, which in fact did grow on trees, were accepted in payment for either goods or labor. Their importance was such that the dishonest buyers and sellers tried to counterfeit cacao by filling the bark or skin of the beans with earth. These fake cacao beans were then adeptly sealed and mixed into piles with true beans, where they could be passed off as genuine. Both the existence of counterfeit beans and the presence of market judges to catch and punish dishonest merchants emphasize the importance of profit as a motive in this precapitalist society.
We know the Aztecs engaged in market exchange. Did other Mesoamerican groups also have markets? Do you think markets may have been present in Mesoamerica before the Aztecs?
y
marketing An exchange system that
frequently involves currencies and generally extends beyond close kinsmen and a small group of trading partners. Market participants try to minimize their costs and maximize their returns to make a profit. cacao A bean of the cacao tree, native
to Mesoamerica; used to make chocolate. Cacao beans also were used as money by the Aztecs.
Concept Aztec Markets
383
_
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 384
Concept
<
Human Sacrifice and Cannibalism Divergent interpretations of ancient rites
Figure 8.71 A depiction of an Aztec ritual heart sacrifice from a sixteenth-century codex.
Many lines of evidence confirm that human sacrifice was one of the most deeply rooted religious traditions of the Aztecs. However, it is clear that the Aztecs were not the only ancient peoples that carried out massacres in honor of their gods, and there is insufficient quantitative information to determine whether the Aztecs were the people who practiced human sacrifice most often. Indeed, sacred texts, literary works, historic documents, and especially evidence contributed by archaeology and physical anthropology, enable religious historians to determine that the practice of human sacrifice was common in most parts of the ancient world. —Alfredo López Austin and Leonardo López Luján (2008)
cannibalism The practice of eating
_ _
human flesh.
384
No issue concerning ancient Mexico has attracted more attention and biting scholarly debate than human sacrifice and cannibalism. The Aztecs have frequently been portrayed in the popular press as violent, insatiable cannibals, practicing bloodthirsty and ghoulish acts (Figure 8.71). Ritual sacrifices have been documented archaeologically in early Mesopotamia, as well as in dynastic Egypt and ancient China. Human sacrifice has a very long history in prehispanic Mesoamerica, perhaps even dating back to the preceramic era in the Tehuacán Valley. Human sacrifice is depicted graphically in several Classic Maya inscriptions and is strongly suggested by the Oaxacan danzante stones. Ancient Mesoamerican societies also gave great ritual importance to blood. Both autosacrifice and the ritual offerings of blood are present in the archae-
Chapter 8 Ancient Mesoamerica
ological and epigraphic evidence (Figure 8.72). Why, then, are the Aztecs so closely linked to the practice of human sacrifice? Most likely it is because the custom was featured so prominently in Spanish descriptions of central Mexico. Whereas the Aztecs believed human sacrifice was a deeply religious practice, necessary for preserving the continuity of the universe, the Spanish Catholics saw it as the devil’s work. Furthermore, the Europeans believed it their duty to abolish this behavior through conquest, conversion, and whatever means possible. The elimination of these customs, which horrified the Spanish, therefore became a rationale for achieving their envisioned destiny to rule New Spain. (It should be noted that, at the time, the Spanish themselves were not above ghoulish behav-
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
ior; in Spain, people were burned at the stake, while in Mexico, many Indians were mistreated and massacred.) Anthropologist William Arens, of the State University of New York at Stony Brook, argues that the Spanish were so interested in defaming the native inhabitants and justifying their conquests that their sixteenth-century accounts cannot be trusted. Arens questions whether cannibalism even existed among the Aztecs. Although the evidence for cannibalism is not as overwhelming as it is for large-scale sacrifice, independent accounts describe the consumption of sacrificed prisoners, particularly by warriors and members of the elite. An alternative to Arens’ explanation has been advanced by Michael Harner, who argues that the absence of domesticated animals and the large number of people in late prehispanic central Mexico necessitated enormous Aztec sacrifices to alleviate a dietary shortage of meat protein and fat through cannibalism. Harner’s ecological interpretation, which lists cannibalism as the reason for enormous sacrifices, has met with little nutritional, economic, or historical support. Harner argues that some 250,000 people were sacrificed each year in central Mexico, whereas most scholars place the annual figure in the tens of thousands. Harner ignores the religious context of Mesoamerican sacrifice and symbolic blood offerings, which often occurred without cannibalism. Many of the Aztec sacrificial events, which elaborated and embellished these prior customs, were dedicatory, carried out as thanksgiving rituals, and involved no cannibalism. The diet of the people of prehispanic Mexico probably was not deficient in either protein or fat. In combination, the dietary staples of maize and beans provide a complete protein source. Such staples were sent in tribute to Tenochtitlán from surrounding polities and distributed through the market system. In one sixteenth-century account, 32 different types of wild fowl, many of which were very fatty,
Page 385
were described as edible; several of these were noted as abundant. Fish and other lake flora and fauna, including high-protein algae that was made into a cheeselike delicacy, also were available, as were domesticated dog and turkey. Some agricultural fields, such as the chinampas, produced two or three crops annually. If, as Harner implies, cannibalism provided the Aztecs with a major source of food, why were most sacrifices held immediately following the major harvest, when food was abundant? In addition, why was human flesh restricted to the warriors and elite, groups likely to have the most to eat anyway? During Cortés’ long siege of Tenochtitlán, the inhabitants were starving, eating almost anything, even adobe and leather. Yet human bodies lay all around, untouched. Aztec human sacrifice is more convincingly interpreted in its historical and symbolic context. The Aztec traditions are connected with earlier Mesoamerican customs of bloodletting, dedicatory offerings, and the sacrifice of war captives in ritual events. The embellishment of the practice of human sacrifice by the Aztecs would seem associated with their distinctly militaristic background. According to Aztec ideology, they were the people chosen to nourish the deities through human sacrifice, as a hedge against uncertainty and to maintain order in the universe. Since sacrificial victims were frequently (though not exclusively) prisoners or slaves, these beliefs may have motivated and fanaticized the Aztec military. The Aztecs conducted several kinds of sacrificial rituals—to bring rain, to commemorate temples, and as a kind of thanksgiving. Large sacrificial events also must have conveyed a powerful message about the strength and sanctity of Tenochtitlán to individuals both inside and outside the city.
Figure 8.72 Two priests pierce their tongues and ears with thorns from the maguey plant in this drawing from a codex.
_ Concept Human Sacrifice and Cannibalism
385
pri31057_ch08.pp320-389.qxd
< ( : < V : y < p : W
10/1/09
Figure 8.73 An example of a ceramic dog figurine with wheels (c. 15 cm tall).
_ _
386
8:12 AM
Page 386
Images and Ideas The End of Prehispanic Civilizations in Mexico The legacy of Mesoamerica’s past
P
eople first set foot in Mesoamerica approximately 10,000–15,000 years ago. From their arrival until A.D. 1519, when the Spanish arrived, Mesoamerican societies developed and diversified independently of contacts with nonnative peoples. Over that period, a Mesoamerican world evolved, composed of diverse groups and polities connected through trade, warfare, intermarriage, and other long-distance contacts and communication. These peoples shared a common cultural background that was distinctively Mesoamerican. Sedentary villages emerged across Mesoamerica during the second millennium B.C., and cities and conquest states were first established in various regions after 500 B.C. The study of ancient Mesoamerica, therefore, enables archaeologists to examine agricultural origins, the emergence of villages and institutionalized social inequality, and the rise and fall of early civilizations. It also permits the investigation of a large macroregion that, unlike portions of the Near East, was never politically controlled by a single empire. Even the powerful Aztec armies were unable to defeat and consolidate the nearby peoples of Tlaxcala or the west Mexican Tarascans. Although longterm change in Mesoamerica did not follow the same course as in Mesopotamia and China (Chapter 10), Europe (Chapter 11), or even South America (Chapter 9), there are many features of these sequences of change and transition that can be compared. The prehistory of Mesoamerica is documented by a variety of sources of information, including both indigenous and sixteenth-century Spanish texts and archaeological studies. With recent breakthroughs in the decipherment of Classic Maya writing systems, scholars studying these lowland societies can now view the past from an indigenous perspective that is not available for other regions of the preconquest Americas. Alternatively, the Mesoamerican highlands are well suited to archaeological surveys. In the valleys of Tehuacán, Mexico, and Oaxaca, as well as in several other highland Mesoamerican zones, the regional perspective from such surveys supplements the more specific findings from careful excavations at individual sites. We have learned that the size and number of prehispanic settlements did not increase at a uniform rate during the prehispanic era and that the pattern of population change was somewhat different in each region. The art and architectural ruins of ancient Mesoamerican societies have captured attention for centuries (Figure 8.73). Now, as we collect more information about their makers, archaeologists have become fascinated with economies, statecraft, religious systems, and other aspects of their lifeways. The diversity of ancient Mesoamerican cities in size and function is truly astounding. Thus Mesoamerica has become a fruitful domain for the study of early urbanism and the processes associated with the rise of ancient cities. Some Mesoamerican cities, such as Teotihuacan, were compact and laid out on a grid plan. Others were dispersed and positioned according to local topography, as in the Maya region during the Classic period. In Classic Maya cities and at Monte Albán, for example, the major thoroughfares were not strictly perpendicular to one another. Residential architecture also was far from uniform at major Mesoamerican cities. Craft manufacturing was central to ancient Teotihuacan, but it played a smaller role in other cities. Although exchange was an integrative mechanism that linked Mesoamerican cities with their hinterlands, marketing may have had an especially crucial role at Tenochtitlán. In ancient Mesoamerica, the volume and magnitude of trade was limited by prehispanic transport. So one can only marvel at the mechanisms that allowed for the organization of tens (and even hundreds) of thousands of people. The longevity of cities such as Teotihuacan and rain-forest Tikal exemplify the effectiveness of the kin, ceremonial, and stately ties that integrated Mesoamerican polities. Such fea-
Chapter 8 Ancient Mesoamerica
pri31057_ch08.pp320-389.qxd
10/15/09
10:59 AM
Page 387
tures as the calendar, the ballgame, certain symbols, sacrificial (blood) rituals, public architectural conventions, and astronomical knowledge were all part of key informational systems that were shared and communicated primarily by the elite, and seemingly central to the maintenance of both their position and societal continuity. Yet the specific ways in which these symbols were used and the nature of Mesoamerican rulership were by no means uniform across time and space. Classic Maya rule emphasized specific lords, individuals like Palenque’s Pakal and Tikal’s Siyaj Chan, who were depicted on stelae in ornate costumes and were eventually buried in elaborate tombs accompanied by rich offerings. In contrast, we currently do not know the names of any of the rulers of ancient Teotihuacan, nor have elaborate burial displays been unearthed. Rule at Teotihuacan seems to have been less focused on specific individuals. Architectural emphasis was placed on large public spaces, such as inside the Ciudadela and the plaza in front of the Pyramid of the Moon. For the most part, mural art depicted ritual feasts and offerings, rather than glorifying the dynastic histories of specific kings. Although Teotihuacan undoubtedly had a ruling class, the strategies of rule and the mechanisms of societal integration were different from those of the Classic Maya. At the same time, whether we look at the Maya region, central Mexico, or Oaxaca, Classic- and Postclassic-period Mesoamerican states went through cycles of growth and decline, centralization and fragmentation. The autonomy of ancient Mesoamerican societies ended in the sixteenth century with the arrival of the Spanish, but the ancestry and culture of the indigenous peoples has in large part survived, helping meld the contemporary lifeways of Mexico and Central America. The foods, artisanry, music, religion, and languages of these areas owe much to their prehispanic heritage. The demise of Aztec society, as well as that of their allies and enemies, has remarkable parallels to similar events farther south. In the northwestern quadrant of South America, the sixteenthcentury Spanish clashed with a powerful indigenous Andean society, the Inca. Inca administration also had come to dominate a large cultural area, even larger than that dominated by the Aztecs. But they, too, were conquered by the arriving conquistadores. In the next chapter, we examine these Andean societies and the series of processes and events that culminated in the empire of the Inca.
p
z
Does this mean that the Maya were then more advanced than their counterparts in, say, Europe? Social scientists flinch at this question, and with good reason. The Olmec, Maya, and other Mesoamerican societies were world pioneers in mathematics and astronomy—but they did not use the wheel. Amazingly, they had invented the wheel but did not employ it for any purpose other than children’s toys. Those looking for a tale of cultural superiority can find it in zero; those looking for failure can find it in the wheel. Neither line of argument is useful, though. What is most important is that by 1000 A.D. Indians had expanded their Neolithic revolutions to create a panoply of diverse civilizations across the hemisphere. —Charles C. Mann (2005)
DISCUSSION QUESTIONS 1. What are some similarities and differences in the settlement plans and residential architecture at Monte Albán, Teotihuacan, and Tikal? What might archaeologists learn from these differences and similarities? 2. How did the Classic Maya centers of Tikal and Palenque differ from later Chichén Itzá? 3. Were all states in indigenous Mesoamerica similar in organization? If so, explain. If not, how were they different?
w w w. m h h e . c o m / p r i c e i p 6 e
For more review material and study questions, see the self-quizzes on your online learning center.
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e. Blanton, R. E., S. A. Kowalewski, G. M. Feinman, and L. M. Finsten. 1993. Ancient Mesoamerica: A comparison of change in three regions, 2d ed. Cambridge: Cambridge University Press. An analysis of long-term change in three Mesoamerican regions. Brumfiel, E. M., and G. M. Feinman, eds. 2008. The Aztec world. New York: Abrams. A collection of essays by renowned Aztec scholars.
Evans, S. T. 2004. Ancient Mexico and Central America: Archaeology and culture history. London: Thames & Hudson. A recent overview of prehispanic Mesoamerica from first occupation through Spanish conquest.
_ Suggested Readings
387
pri31057_ch08.pp320-389.qxd
10/15/09
10:59 AM
Page 388
Flannery, K. V., ed. 1976. The early Mesoamerican village. New York: Academic Press. A humorous, yet highly instructive, volume on ancient Mesoamerica that illustrates how archaeology should proceed at various scales of analysis to interpret the past. Flannery, K. V., and J. Marcus, eds. 1983. The cloud people: Divergent evolution of the Zapotec and Mixtec civilizations.
_ _
388
Chapter 8 Ancient Mesoamerica
New York: Academic Press. Reprint, Clinton Corners, NY: Percheron Press, 2003. The synthetic overview for prehispanic Oaxaca. Sharer, R. J., and L. P. Traxler. 2006. The ancient Maya, 6th ed. Stanford, CA: Stanford University Press. A classic text on the Maya.
pri31057_ch08.pp320-389.qxd
10/1/09
8:12 AM
Page 389
_
pri31057_ch09.pp390-435.qxd
10/1/09
C
H
8:16 AM
A
Page 390
P
9
T
E
R
N
I
N
E
Figure 9.1 Gold bead in the form of a human head, 5.3 cm (2 in) high, from Tomb 3 at Sipán. This bead was part of a necklace (10 beads in all) that was placed near the neck of the principal individual in the tomb.
_ _
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 391
South America The Inca and Their Predecessors
Introduction Prehispanic South America The foundations and cycles of Andean states and empires
I
n Peru (as in too many other places in the world), the tradition of pillaging ancient sites for priceless artifacts is an old one, extending back to Spanish conquerors, who, beginning in 1532, ransacked the Inca empire for gold, silver, and precious stones. Around midnight one evening in February 1987, Peruvian archaeologist Walter Alva, director of the Brüning Archaeological Museum in Lambayeque, Peru, was called by the local chief of police, who reported with urgency that there was something that Alva had to see. Looters had illegally tunneled into an earthen pyramid built by the ancient Moche (A.D. 100–700) at Sipán and exposed burials with associated metal, shell, and ceramic artifacts that the grave robbers intended for the international art market. In response to the looting, Alva and his wife, Susana Meneses, began years of careful salvage excavations in this earthen mound at Sipán, scientifically investigating the richest tombs ever found in the Americas (Figure 9.1). Their excavations have made sense of these fabulous finds, placing them in a cultural context that could never have been achieved if the entire mound had been pillaged. In this chapter, the antecedents of the Moche, as well as the web of history that ran from the Moche to the Inca, are explored for this ancient Andean region. When Christopher Columbus left Spain on the voyage that brought him to America, the largest empire in the world was Tawantinsuyu, the Inca empire. Much greater in size than any fifteenth- or twentieth-century European state, Tawantinsuyu (“land of four quarters”) stretched over 984,000 sq km (380,000 sq mi), about the size of Washington, Oregon, Idaho, and Montana combined. The Inca controlled the most extensive political domain that has ever existed in the Southern Hemisphere. Cuzco, the capital of the realm, governed some 80 provinces that sprawled from Colombia in the north across the Andean highlands of Argentina, Bolivia, Peru, and Ecuador to Chile in the south. From the Pacific Ocean, the Andes rise so abruptly that a mere sliver of flat coastal land separates the high mountains from the water. This coastal strip is one of the world’s driest deserts. In the high Andes, where peaks rise to more than 6100 m (20,000 feet) above sea level (the highest peaks in North or South America), humans occupied the relatively flat basins and valleys, as well as the high grassland plateaus (punas). The eastern side of the mountains, called the montaña, is wet and heavily vegetated. Farther east lies the enormous drainage basin of the Amazon, composed of tropical forests and savannas. Feathers and many tropical plants were brought from the forest into the Andean and coastal zones.
391
< ( : < V : y < p : W
CHAPTER OUTLINE
Introduction Prehispanic South America 391 SITE El Paraíso
395
CONCEPT The Maritime Hypothesis 399 SITE Chavín de Huántar CONCEPT The Textiles of Paracas 404 SITE Moche
405
CONCEPT The Nazca Geoglyphs 410 SITE Sipán
411
SITE Tiwanaku SITE Chan Chan
415 419
SITE Cuzco and Machu Picchu 423 CONCEPT Inca Highways SITE Huánuco Pampa
400
427
429
Images and Ideas
The Organization of State Society 433
_
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
w w w. m h h e . c o m / p r i c e i p 6 e
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
The myth persists that in 1492 the Americas were a sparsely populated wilderness, “a world of barely perceptible human disturbance.” There is substantial evidence, however, that the Native American landscape of the early sixteenth century was a humanized landscape almost everywhere. Populations were large. Forest composition had been modified, grasslands had been created, wildlife disrupted, and erosion was severe in places. Earthworks, roads, fields, and settlements were ubiquitous. With Indian depopulation in the wake of Old World disease, the environment recovered in many areas. A good argument can be made that the human presence was less visible in 1750 than it was in 1492. —William M. Denevan (1992)
_ _
392
Page 392
Our main focus in this chapter is Andean South America. The people of the Amazon rain forests and savannas had a unique and important prehistory as well. Yet because of the impenetrability of the lush environment, archaeologists and ethnohistorians still know much less about the latter area than about the more accessible coastal and mountain zones. In 1948 and 1949, Betty Meggers and Clifford Evans, of the Smithsonian Institution, began preliminary investigations on Marajó Island at the mouth of the Amazon River. On this large island, 39,000 sq m (15,000 sq mi), huge earthworks were found, which Meggers and Evans interpreted as the remnants of a complex society that had moved to the island from the Andes, only to meet a rapid demise. At the time, this interpretation was well reasoned because the large scale of the earthworks was unexpected, given their absence in small, egalitarian, contemporary Amazonian Indian villages. Archaeological work on Marajó Island by Anna Roosevelt, of the University of Illinois at Chicago, and her colleagues has revised our perspective on Marajoara origins. Their research illustrates numerous continuities between modern Amazonian people and the earlier Marajoara occupation and documents a more-than-1000-year habitation sequence (400 B.C.–A.D. 1300) on the island. For example, Roosevelt found large, globular burial urns that were painted to represent crouching humans, decorated with multicolored, geometric patterns similar to the abstract drug visions illustrated by contemporary Amazonians. Marajoara art also has antecedents in earlier lowland styles, and the painting method is Amazonian (not Andean) in character. The physical affiliation of Marajoara skeletons is closer to Amazonian than to Andean people. At the same time, reanalysis of written accounts of sixteenth-century Spanish penetration into the Amazon, along with recent archaeological mapping and excavations, has led researchers to argue that many of the late prehistoric populations of this region lived in more permanent settlements that were more densely distributed than found for more contemporary settlements. The past settlements also were more hierarchically organized. Differences, such as the cessation of inherited positions of leadership, between pre-Columbian inhabitants and contemporary groups may owe much to the diseases brought by the Europeans. Western South America contained the continent’s first agricultural settlements (e.g., Guitarrero Cave), as well as its earliest sedentary communities. We examine the prehispanic peoples of this area, beginning with early monumental construction on the Peruvian coast and neighboring river valleys. We start with the early site of El Paraíso, a sedentary coastal community dating to before 2000 B.C., and conclude with the late prehispanic Inca sites of Cuzco, Machu Picchu, and Huánuco Pampa (Figure 9.2). The prehistory of this part of South America is, in a general way, similar to that of Mesoamerica. In both areas, the earliest experiments with food production preceded the transition to sedentary village life. In South America, the first villages date to 3500 B.C., much earlier than in Mesoamerica. In both regions, however, these first sedentary communities were situated near the coast. Almost immediately, greater social differences and the first positions of leadership become evident in the archaeological record. In Mesoamerica, these changes occurred along with the development of an interregional style, the Olmec Horizon, represented in exotic and highly crafted items found in both the highlands and the lowlands. Between 900 and 200 B.C., a similar phenomenon, the Chavín style, was shared from the edges of the Amazon to the Pacific Coast. This style is named for the carvings found on a temple in the uplands of central Peru at Chavín de Huántar. In Peru, however, the spread of the Chavín style postdates the construction of truly monumental public architecture and the emergence of social differences at coastal sites that were contemporary with earlier El Paraíso. In South America, the end of the last century B.C. was marked by the rise of major centers that became the core of states administering regional populations. We discuss Moche, on the north coast, and refer to the fantastic line drawings
Chapter 9 South America: The Inca and Their Predecessors
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 393
ATLANTIC
PACIFIC OCEAN
OCEAN
COLOMBIA
Equator
Marajó Island
ECUADOR
Amazon
Sipán
Pampa Grande Galindo Moche Chan Chan Chavín de Huántar Pañamarca Huánuco Pampa Aspero Caral PERU El Paraíso Wari Machu Picchu Cerro Azul Cuzco Paracas Cahuachi
BRAZIL
Lake Titicaca
Tiwanaku Chiripa BOLIVIA
Sites emphasized in text Other sites mentioned in text Modern international boundary
CHILE 0
ARGENTINA
300 km
El Paraíso Chavín de Huántar Moche Sipán site at its largest extent
Tiwanaku Chan Chan Cuzco Machu Picchu
episode of occupation
2000
1500
Huánuco Pampa 1000
500
B.C./A.D.
500
1000
Year
constructed by the Nazca in the south. We also discuss the coastal site of Sipán, where spectacular Moche-era tombs were excavated. Later, Tiwanaku, a giant economic and religious center near the southern end of Lake Titicaca in Bolivia, began to incorporate larger areas outside its local region. As with the Aztecs in the Central Highlands of Mexico, the end of the prehispanic era in South America was characterized by a powerful people, the Inca, whose rulers exacted tribute from a wide domain. Yet this much larger South American empire had clearer imperial predecessors, such as the Tiwanaku polity and the Chimú kingdom. After A.D. 1000, the latter, which was centered on the north coast at Chan Chan, consolidated a large, primarily coastal domain. This political sphere was eventually engulfed by the expansive Inca, who conquered the Chimú between A.D. 1462 and 1470. We examine the nature and organization of this Western Hemisphere empire by looking at Cuzco, the Inca capital, the
1500
Present
Figure 9.2 Location of and timeline for South American sites.
Introduction Prehispanic South America
_
393
pri31057_ch09.pp390-435.qxd
10/1/09
quipu The Inca word for an elaborate
knotted string device used by the Inca and other peoples in Peru for record keeping. binary-coded A system of informa-
_ _
tion storage or processing in two states (such as 0 and 1). Computers process immense amounts of information through the manipulation of such sequences.
394
8:16 AM
Page 394
famous high mountain settlement of Machu Picchu, and the provincial administrative center of Huánuco Pampa. Although parallels can be drawn between the long prehispanic sequences in Mexico and Peru, there also were important differences. Most of the major Peruvian centers were shorter-lived than Mesoamerican sites such as San José Mogote, Monte Albán, and Teotihuacan, which were inhabited almost continuously for a millennium or more. The Andean world also seems to have lacked a core region like the Basin of Mexico, which dominated the larger landscape for much of the prehispanic era. In South America, the balance of power seems to have shifted repeatedly between the Pacific Coast and the rugged uplands. Likewise, at times, the core of the Andean world was positioned in the south (as at Inca Cuzco); at other times, the Andean world’s most powerful centers lay in the north at Moche and Chan Chan. Animal domestication was more important in South America, where a variety of camelids, in addition to guinea pigs, were tamed. Since only the turkey, the dog, and the honeybee were domesticated in Mesoamerica, long-distance land transport was conducted entirely by human bearers. In South America, the llama (a camelid), used also for wool and meat, served as a pack animal for loads up to 45 kg (100 lb), making land transportation considerably less costly than it was in Mesoamerica. Llamas, like the more efficient Eurasian beasts of burden, subsist on grasses and other foods that humans cannot digest. The Inca built and maintained an extensive, often paved road system (with suspension cables, stone bridges, and a system of relay runners) that crossed almost the entire empire and was unmatched in Mesoamerica. Watercrafts, including giant trading canoes, were more developed in Middle America. In Mesoamerica, a writing system developed more than 2000 years ago. Subsequently, it diversified, with several systems, particularly that of the Maya, becoming more complex (with relatively long texts). In contrast, to date no archaeological evidence for a traditional form of writing has been found in the prehispanic Andes. The Inca did possess an ingenious numerical apparatus, the quipu (knot), which was composed of a main horizontal cord from which a series of smaller strings hung. The quipu was used to record numbers, by tying knots at various intervals on these strings. The knots farthest from the main cord represent the smallest numbers. Recent studies of the quipu have proposed that it was used by the Inca to convey more information than just complex mathematical data. The color, placement, and nature of the various knots on quipu are now thought to have been used to record and transmit more intricate messages and information through a system of binary-coded sequences. Even without a traditional system of writing, the Inca were able to establish a larger and better integrated political system than ever existed in prehispanic Mesoamerica. Although the written words left by the Maya and other Mesoamerican peoples are absent in the Andes, archaeologists working along the arid coast of Peru have the advantage of a quality of preservation unknown almost anywhere else. Standing adobe architecture frequently sits almost unscathed by the elements for centuries, even millennia. The recovery of large woven textiles is not unusual. Preservation can be so good that even fingernail clippings and human hair were found by Joyce Marcus using careful excavation procedures at coastal Cerro Azul, a site that postdates A.D. 1100. Marcus also found a series of mummified burials at Cerro Azul. Males generally were buried with fishing nets or slings (for hunting), and females were buried with an array of different weaving implements (including camelid bone tools, bobbins made of thorn, needles, and balls of various colored yarns). South America provides another fascinating area for studying the emergence of an indigenous civilization. Only through archaeology can we hope to discover the remarkable similarities—as well as the all-important differences— that characterized these long cultural processes in various regions of the globe.
Chapter 9 South America: The Inca and Their Predecessors
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 395
Site
:
El Paraíso Settlement and monumentality on Peru’s coast and adjacent river valleys
The waters off the west coast of Peru are one of the world’s richest fishing areas. A complex mixing of ocean currents provides an extraordinary source of marine foods. The shore of these abundant waters, however, is one of the world’s driest deserts. The same cold waters that foster rich plankton and fish life also inhibit rainfall on the coast. This coastal zone, stretching from the shoreline to the foothills of the Andes, only rarely receives measurable quantities of precipitation. Short-lived torrential downpours occur only once or twice a decade. Rivers and streams carrying the snowmelt and rainfall from the Andes provide most of the surface water, cross-cutting the bleak coastal desert in an east-west direction. These waterways are in essence isolated verdant pockets along the arid coastline. The desert coast of central Peru was first settled after 7000 B.C. by mobile groups who exploited a wide range of environmental zones. The Pacific Ocean was a source of fish and shellfish; deer, small mammals, and birds were hunted and wild plants collected in the coastal river valleys. The dry foothills of the western Andes become oases of scattered communities of vegetation, or lomas, between June and October (winter in the Southern Hemisphere). Here the early inhabitants ate deer and the edible seeds of grasses or sagelike plants. Grinding stones used in preparing these plant foods have been found in these areas. After 5000 B.C., certain coastal populations relied increasingly on marine and plant products, including cultivated squash and tubers that had been introduced to the desert coast from the highlands. Peru’s first permanent, year-round villages were established after 4000 B.C., more than 1000 years before the first Peruvian evidence for pottery.
Based on recent research by Ruth Shady Solis, of the National University of San Marcos, Peru, preceramic sedentary settlements with monumental architecture were established around 3000 B.C. in several river valleys just inland from the coast. Caral is the largest and best known of these centers (Figure 9.3). Located in the Supe Valley, the settlement consists of a complex of six platform mounds, plazas, and residential buildings that cover 60 ha (150 acres). The large mounds were raised above the surrounding desert through a construction technique of placing shicra in the fill. Inside the ring of pyramids is a large sunken amphitheater. Regional surveys and test excavations by Jonathan Haas, of the Field Museum of Natural History in Chicago, and Winifred Creamer, of the University of Northern Illinois, have revealed other monumental settlements in the Supe and adjacent river valleys along a stretch of the Peruvian coast known as Norte Chico. Other centers, perhaps slightly later in date, also were established on the coast. One of these sites, Aspero, at
Figure 9.3 Main complex of platform mounds at Caral in the Supe Valley.
lomas (Spanish) Vegetation that is supported by fog in otherwise arid environments. shicra (Inca) Meshed bags containing
rocks, used as fill in the construction of ancient Andean structures.
Site El Paraíso
395
_
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 396
Figure 9.4 An artist’s reconstruction of architecture at Aspero. The largest mound at this 13-ha (32-acre) site stands over 10 m (33 ft) high.
_ _
Figure 9.5 A fragment of cotton net used for fishing.
396
Chapter 9 South America: The Inca and Their Predecessors
the mouth of the Supe River, has been excavated by Robert A. Feldman (then at the Field Museum). Aspero consisted of multiroomed compounds of interconnected rooms of various sizes that were divided by stone walls (Figure 9.4). The outer rooms were usually the largest and had the widest doorways. The inner rooms, which contained cached artifacts and wall niches, were small. These central rooms could be entered only by winding through the more open larger rooms and then stepping through narrow entryways. Thus access to the inner cores of the nonresidential structures may have been restricted to a small segment of the population. In contrast to the coastal sites, inland valley centers were located near sources of water that could be tapped for irrigation for growing a variety of plants such as squash, sweet potatoes, and beans. But surprisingly, cotton, and not edible plants, was the most abundant crop at Caral. Remains of seafood, especially sardines and anchovies that must have come from the coast, also were recovered in abundance at Caral. Cotton fishing nets have been found at Aspero (Figure 9.5). The exchange of cotton from inland sites to coastal people to make nets for fishing while inland sites received sea resources from coastal villages indicates that a symbiotic relationship existed between the coastal and inland sites. After 2500 B.C., larger settlements with monumental architecture were established along the coast. El Paraíso (EL par-a-EASE-o) one of the largest and better studied of these preceramic settlements, was located about 2 km (1.2 mi) from the coast along a permanent stream, next to a wide area of irrigable land. El Paraíso contains eight or nine large stone structures that, in total, cover over 58 ha (140 acres) (Figure 9.6). The structures range in size from three or four rooms to enormous complexes 300 m (980 ft) long ⫻ 100 m (325 ft) wide. The site represents an early example of monumental construction and nucleated population in the Americas. Although El Paraíso has been known to archaeologists for decades, it
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 397
VII
PERU ón R iv
er
X
N
Chill
El Paraíso
II
IX N III
VI
VIII
IV
V XI
Hills Water course Architectural unit
I 0
100 m
Figure 9.7 The plan of Unit I, a partially restored architectural complex in the southwestern part of El Paraíso.
Figure 9.6 The plan of El Paraíso, whose name translates as “paradise.” The village was situated near the mouth of the Chillón River on the central coast of Peru.
was not excavated until 1965 by Frederic Engel. The first excavations indicated an early preceramic date and revealed that one of the stone structures (Unit I) consisted of a series of rectangular rooms, courts, and passageways delineated by walls of 1.5–2.5 m (5–8 ft) in height (Figure 9.7). The walls were built of large stone blocks cemented together with clay and faced with a mud plaster painted with black, red, ochre, and white. The compound was filled in with rock rubble and then rebuilt several times during the occupation of the site. Excavations by Jeffrey Quilter and his colleagues have confirmed that two other compounds also were preceramic. Radiocarbon dates from the three compounds indicate a short, possibly 300-year contemporaneous occupation that ended by 1500 B.C.
Despite these findings, many questions remain. How were these masonry compounds used? Although some of the rooms in Unit I contain possible offerings and may have been used for ceremonies, the diversity of artifacts (grinding stones, a wooden sling, stone and wooden tools) from the compound suggests it was not simply a religious structure. Quilter found domesticated cotton, wool, and needles in one room in Unit II, suggesting that this area may have been used for making textiles. Large quantities of guano, a bird offering, and colorful bird feathers also were found in the same room, likely evidence for the ancient practice of weaving colored feathers into high-status cloths. Other cotton was woven into nets and lines used in fishing. Although there is little
How does the sequence of transition to agriculture, sedentary life, and pottery compare between Middle America and the Andean region?
y
guano Bird excrement.
Site El Paraíso
_
397
pri31057_ch09.pp390-435.qxd
10/1/09
Figure 9.8 The stepped stone façade of the north face of Unit I at El Paraíso.
_ _
coprolite Fossilized feces.
398
8:16 AM
Page 398
consensus about the kind of sociopolitical organization at El Paraíso, large quantities of labor clearly were required to quarry and shape the stone blocks used to build the compounds (Figure 9.8). Additional coordination, planning, and muscle power were needed to repeatedly rebuild the structures. Yet the biggest mystery concerns the diet of the residents of El Paraíso and the other preceramic inhabitants of coastal Peru. In 1975, Michael Moseley, of the University of Florida, proposed that the foundations of later Andean civilization were established during the preceramic era at such places as El Paraíso. He believed that the inhabitants depended largely on marine foods. Although no one disputes the important role of foods from the sea (shellfish, marine mammals, and fish), various plants, including the starchy tropical tubers (jicama and achira), cultivated common beans and lima beans, and domesticated chili peppers, gourds, and squash have all been recovered at El Paraíso, along
Chapter 9 South America: The Inca and Their Predecessors
with the remains of birds and small quantities of land mammals. More than 90% of the faunal remains identified by Quilter and his team belonged to bony fish and mollusks. However, the site’s location, adjacent to an unusually large patch of arable land near the mouth of the Chillón River and 2 km (1.2 mi) from the Pacific, also indicates a terrestrial component important to the preceramic economy (cotton production) and diet. This inference is supported by the analysis of ten human coprolites from El Paraíso, which revealed that both squash and a variety of wild plant foods (including grasses, seeds, and tree fruits) constituted a significant portion of the diet. Many questions remain regarding the early inhabitants of El Paraíso and other preceramic settlements in coastal Peru. Yet there is no doubt that the archaeological work at this site and others along the dry Pacific Coast has revolutionized our knowledge of ancient South America and its early sedentary beginnings, thousands of years ago.
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 399
Concept
<
The Maritime Hypothesis Were the roots of Andean civilization on the Pacific Coast?
In 1975, Michael Moseley proposed that complex societies with monumental architecture emerged during the preceramic period on the Pacific Coast of South America. Subsistence activities at these sites appear to have focused on marine resources. Moseley intended to refute the widely accepted belief that only agricultural economies can support the foundations of civilization. He bolstered his argument, which he called the Maritime Hypothesis, by noting the abundance of shellfish middens and fishing nets at many coastal settlements, particularly at several sites that were closer to the coast and somewhat earlier than El Paraíso. Although there is little doubt that marine foods were an essential component in the diet of the prehispanic coastal dwellers, several studies have suggested that both wild and agricultural plants also were a significant part of ancient Peruvian subsistence. Root and seed crops have been recovered in relatively small quantities in midden deposits at El Paraíso, as well as at other preceramic coastal sites. However, the quantity of plant remains, compared with the abundance of shell, may not provide an accurate reflection of the relative importance of those foods in the ancient diet. Plant remains generally do not preserve in the archaeological record nearly as well as harder materials, such as mollusk shells. Poor preservation is a particular problem for detecting the root crops that were important in ancient Peru, because they lack the harder parts that make seed crops somewhat easier to find. The problems of preservation were aggravated in many early coastal excavations where fine screening and other meticulous recovery techniques were not used. Several critics of the Maritime Hypothesis have noted that a subsistence
regime largely dependent on the abundant marine resources of the Peruvian Pacific would have left the coastal inhabitants at the mercy of the warm countercurrents that periodically disrupt the bounty of the ocean environment (Figure 9.9). Five to ten times each century, a countercurrent, referred to as El Niño (“the Christ Child”), moves in from the north soon after Christmas, altering the normal patterns of water temperature, flow, and salinity, and diminishing the availability of nutrients. El Niños generally last 2–12 months and severely disturb marine life. Bountiful schools of fish and flocks of seabirds either migrate or die. Because of the irregularity of these maritime disruptions and the difficulty of storing large quantities of marine products in a warm climate, it seems likely that plant foods provided a significant portion of the ancient diet. In fact, this suggestion has been supported by Jeffrey Quilter’s recovery at El Paraíso of a diversity of wild and domesticated plant foods, through both coprolite analysis and the sifting of excavated trash deposits through 1⁄4-inch hardware cloth and 1⁄16-inch mesh. The recent discoveries that inland communities like Caral may have been earlier, larger, and more monumental than coastal settlements also has prompted questions concerning some of the expectations of the Maritime Hypothesis as originally proposed. Nevertheless, current research places the early roots of Andean civilization in this broadly defined coastal setting where the Andes grade down into the Pacific.
Figure 9.9 A drawing of ocean currents and coastal weather patterns. Moisture-bearing clouds moving from the Pacific Ocean are warmed by the landmass, but temperature inversions keep precipitation from falling until the higher altitudes of the Andes are reached (about 2500 m), leaving the coast dry but cool. The waters off Peru are swept northward by the coastal current. The deeper undercurrent flows southward; upwelling waters from the lower depths bring up rich nutrients that support abundant marine life.
El Niño (Spanish) A warm-water
countercurrent that periodically appears off the Peruvian coast, usually soon after Christmas, and alters the normal patterns of water temperature, flow, and salinity.
Concept The Maritime Hypothesis
399
_
pri31057_ch09.pp390-435.qxd
10/15/09
11:01 AM
Page 400
Site
:
Chavín de Huántar An early Andean center
oca A South American wood sorrel
(Oxalis crenata) that is cultivated for its edible tuber. quinoa A pigweed (Chenopodium
_ _
quinoa) of the high Andes. Seeds of the plant were ground and used as food in the past and still are today.
400
Figure 9.10 The ruins of the Castillo, the largest platform at Chavín de Huántar, situated at the base of the Cordillera Blanca, between the Huachecsa and Mosna rivers.
Chapter 9 South America: The Inca and Their Predecessors
The Peruvian coast and neighboring river valleys, homeland for such late prehispanic polities as the Nazca, the Moche, and the Chimú, also gave rise (before 2000 B.C.) to the first complex political systems in western South America. Centralized polities, focused on large sites that contained monumental architecture, including ceremonial mounds and plazas, developed in a number of valleys along the north and central coasts. After 1000 B.C., the balance of power began to shift upward as larger centers developed in the Andean highlands. Perhaps the best known of these sites is Chavín de Huántar (cha-VEEN day WHAN-tar), situated more than 3000 m (10,000 ft) above sea level in what is today a typical Peruvian valley (Figure 9.10). The site is surrounded by steep slopes, including the snowcapped Cordillera Blanca, whose highest peaks rise above 5500 m (18,000 ft). Chavín de Huántar is located in terrain that was well suited to a mixed agricultural strategy of herding camelids and farming high-altitude tubers (potato, oca) and native grains (quinoa). Natural grasslands provided rich habitat for camelids, deer, and other wild game while glacial lakes and snow from the mountains provided supplemental water for irrigating the narrow valley floor. Other food items that could not be produced locally, such as chili peppers, salt, and dried fish, were obtained through exchange. Located near the confluence of two rivers, Chavín de Huántar was ideally situated to control natural transportation routes to the west (toward the coast) and into the vast Amazonian lowlands to the east. In the early years of the Colonial era, Spanish chroniclers noted the grandeur of the ruins and the beautiful decoration visible in the finely cut stones at Chavín de Huántar (Figure 9.11).
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Yet the historical placement of the ruins and their importance to Andean prehistory were established over 60 years ago, through the works of the Peruvian archaeologist Julio Tello. Tello correctly recognized the important artistic similarities between the stone carvings at Chavín de Huántar and the decorative items of pottery, stone, and metal found at other sites in the highlands and on the coast. These similar motifs define a stylistic pattern that has become known as the Chavín Horizon. Typically, Chavín carvings interweave figures that combine the natural features of people, snakes, jaguars, caymans, and birds with intricate geometric and curvilinear designs. Radiocarbon dates from Chavín de Huántar by Richard L. Burger, of Yale University, place the major occupation of the site between 850 and 200 B.C. It appears that construction of the civic-ceremonial architecture began early in the occupation but did not reach its largest (more than 40 ha, or 100 acres) and most monumental extent until after 390 B.C. By 200 B.C., the central core of the site included a complex of rectangular, stone masonry platforms covering an area greater than 4000 sq m (about 1 acre). Although the height of this construction is difficult to assess because the main structures were on a slope, the largest platform, the Castillo, or New Temple, rises about 13 m (45 ft) above the surrounding terrace. The Castillo platform was not solid masonry but was composed of three superimposed levels of small, interlocking galleries and rooms, which were covered by flat slabs and separated by adobe and stone walls. The interior walls of these mazelike passageways were sometimes painted, and the inside rooms were ventilated by a complex network of shafts. The platform exterior was faced with dressed granite, and at one time, these outer walls were adorned with sculpted human and animal heads (Figure 9.12). The earlier so-called Old Temple at Chavín, under the northwest corner of the New Temple, was U-shaped and consisted of a main building with two
Page 401
Figure 9.11 This finely carved stone from Chavín de Huántar, measuring almost 2 m (6.6 ft) high, depicts a deity holding two staffs and wearing an elaborate headdress.
wings that formed an open square court (see Figure 9.13). In the center of this open area was a sunken circular plaza. Cut stone staircases on the east and west provided access to this semisubterranean feature (Figure 9.14). The Old Temple and the rectangular court surrounding the circular plaza also were honeycombed with a series of interior passageways and galleries. Deep inside the Old Temple, in a chamber under the central, eastern part of the U-shaped platform, stands a carved prismlike shaft of white granite. This tall pillar, named the Great Image or
The construction of monuments, such as pyramids or massive ceremonial mounds, and the rearrangement of the landscape, such as the construction of artificial hills or barrows, require considerable labor and resources. In large measure, it is difficult to tell a lie with monuments. Yes, a monument will be placed on a hill to make it more imposing, but observers learn quickly to discount such tricks, to look behind the facades. Monumental constructions require leadership, coordination, and finance. They are inherently expensive in terms of a group’s resources, requiring many people to work together for long hours. —Timothy Earle (1997)
cayman A tropical South American
alligator.
Site Chavín de Huántar
_
401
pri31057_ch09.pp390-435.qxd
y
10/1/09
8:16 AM
Page 402
What behaviors might lie behind the phenomenon that archaeologists term the Chavín Horizon?
Figure 9.12 Sculpted stone head with mix of human and animal features.
Lanzón, was placed in position when the Old Temple was built (Figure 9.15). Its pointed base was set deeply into the chamber’s floor, and its top was fitted into a space in the gallery ceiling. Each side of the Great Image was carved in bas-relief, depicting a standing human
figure with feline teeth and nostrils. The anthropomorphic figure was wearing earplugs, a necklace, and a belt that contained serpent-jaguar faces. These images also adorned the upper portion of the monolith, as a kind of headpiece perched on the serpent hair of the main figure. The Great Image was just one of a number of large stelae found at the site. By 400–200 B.C., several thousand people lived at Chavín de Huántar, with residential areas surrounding the ceremonial complex. In size and scale of construction, the site far surpassed any other occupation in the narrow Mosna Valley, where it was located. During this period, Chavín de Huántar also became much more interconnected with large settlements in other highlands, as well as on the coast. Although some archaeologists have proposed that the sharing of the Chavín style reflects some kind of broad, centralized religious or economic control, this now seems highly improbable. Rather, this wide distribution of cer-
Figure 9.13 Plan of the Old and New Temples at Chavín de Huántar.
NEW TEMPLE
OLD TEMPLE
“CASTILLO” Lanzón O L D
T E M P L E
Dual summit buildings
Circular plaza NEW TEMPLE
Rectangular sunken court
N 0
30 m
Gallery
_ _
402
Chapter 9 South America: The Inca and Their Predecessors
20 m 0
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 403
Figure 9.14 Staircase of cut-stone blocks on the west side of the sunken circular plaza.
tain ideological and stylistic conventions may mark an intensified level of exchange, communication, and militaristic competition between highland and coastal settlements. In certain ways, the Chavín Horizon is similar to the Olmec Horizon in Mesoamerica, although unlike with the latter, sociopolitical differences were present in certain South American polities for centuries before the spread of the Chavín style.
Figure 9.15 The Lanzón, deep inside the Old Temple of Chavín, stands 4.5 m (15 ft) high and weighs about 2 tons.
_ Site Chavín de Huántar
403
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 404
Concept
<
The Textiles of Paracas Clues to visual communication in ancient South America
Figure 9.16 A cloth mantle of the Paracas style. Images on ritual weavings played a key role in the visual communication of prehispanic Peruvian society, particularly at the south coastal sites associated with this style. The anthropomorphic figure shown in this weaving was a common image depicted on Paracas textiles.
The craft of using fiber to produce goods, ranging from floor mats to bridal dowry, addresses all three of the basic human needs: food, clothing, and shelter. Cordage, basketry, and matting, as well as woven cloth, have long been such an integral part of human adaptations that one can hardly pass through a day without using a metaphor that is ultimately derived from the production of fiber-related goods. —Irene Good (2001)
shaman An anthropological term for a spiritualist, curer, or seer.
_ _
repoussé (French) The process of forming a raised design on a thin sheet of metal by placing it over a mold and hammering it in place.
404
In ancient Peru, textiles were an important means of visual communication. This use of woven cloth to convey ritual images is perhaps best exemplified on the south coast of Peru by the Paracas (par-RAH-kas) culture (about 600–150 B.C.). Paracas elite wore lavish costumes of mantles, ponchos, skirts, tunics, loincloths, and woven headgear. Even after death, these individuals were covered with many fine weavings. On the dry south coast, careful excavations have uncovered mummified burials still wrapped with brightly colored, embroidered ritual cloths, providing a vivid picture of the symbolic life of these ancient Peruvians. Although the Paracas textiles have their own distinctive style (Figure 9.16), they do share certain artistic features with contemporary Chavín cloths. Both portray anthropomorphic and composite zoomorphic figures. Yet on the brightly colored Paracas textiles, these figures have a less somber presence than the supernatural creatures depicted in stone and pottery at Chavín de Huántar. One of the most frequent images found on the Paracas textiles is a human
Chapter 9 South America: The Inca and Their Predecessors
figure that appears to be either a shaman or a person impersonating a supernatural figure. Some of these individuals are elaborately costumed, often dressed like the individuals whose bodies the textiles bundle. Arched sharply backward with outstretched arms, head askew, and flailing hair, these figures seem frozen in the midst of rapid movement, perhaps magical flight. The Paracas textiles provide a dramatic record of ancient Peruvian weaving skills. Copper and gold ornaments attest to the metallurgical proficiency of other South American peoples during the same period. Small copper tools and ornaments have been recovered in northern Bolivia, where the cold metal was worked by hammering and shaping. In the north, gold was made into earspools, as well as nose, chest, and other head adornments. The latter were crafted by working the metal into thin sheets and then hammering it over a mold, a process referred to as repoussé. These gold ornaments, which also frequently were soldered or welded, often bore the composite figures characteristic of the Chavín Horizon.
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 405
Site
:
Moche Giant pyramids on Peru’s north coast
che
Rive
r
Modern fields
Mo
The end of the Chavín phenomenon marked the start of an era, 200 B.C.– A.D. 600, that was primarily characterized by the development of small local polities and their associated distinct regional ceramic traditions. Around the time of Christ, the site of Moche (MO-chay) was established about 6 km (3.5 mi) from the ocean near the modern city of Trujillo in the Moche Valley, on the north coast of Peru. The site is located 600 m (2000 ft) from the Moche River, close to the center of the Moche Valley, and up against Cerro Blanco, a large hill that rises abruptly to a height more than 500 m (1650 ft) above sea level. In its day, Moche was the largest settlement on the north coast of Peru. Unlike most other centers of its time, Moche controlled not only the rest of the Moche Valley but also neighboring areas for centuries. The Moche site is dominated by two major pyramids, or huacas: the Huaca del Sol (Figure 9.17) and the Huaca de la Luna. The two structures are separated by a 500-m (1650-ft)-wide plaza or level area. The Huaca del Sol towers 40 m (130 ft) above the surrounding plain and measures at least 340 ⫻ 160 m (1100 ⫻ 525 ft, larger than a city block) (Figure 9.18). It was one of the largest, if not the most massive, solid adobe structures ever built in the Americas. The pyramid once contained more than 130 million sun-dried bricks. The Huaca de la Luna, which rests on the lower slopes of Cerro Blanco, is a 30-m (100-ft)-high complex of platforms, walls, and courtyards. Made of adobe, it was formed by three mounds connected by courts and was constructed in at least three stages. The two pyramids were used differently. Activities on top of the Huaca del Sol were associated with the accumulation of domestic refuse, whereas the summit of the Huaca de la Luna
Cerro Blanco
Huaca del Sol Huaca de la Luna
N
0
200 m
was swept clean, and the uppermost structures were decorated with painted murals. The murals were found in 1899 by the archaeologist Max Uhle, who also discovered a cemetery area between the two pyramids. More recently, Michael Moseley and his colleagues have studied the adobe bricks from the two huacas and have found more than 100 different symbols impressed on their top surfaces. Because these symbols were applied during production, Moseley reasoned that they may have been makers’ marks (stylized signs), used to distinguish the bricks manufactured by different groups (Figure 9.19). The two huacas, which required hundreds of
Figure 9.17 The plan of Moche. Bounded by Cerro Blanco on one side and by the Moche River on the other, the two large huacas are separated by a large open area or plaza that measures half a kilometer across.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on Moche art and other aspects of this ancient society, see the Internet exercises on your online learning center.
huaca (Quechua) An Andean word
for pyramid.
Site Moche
_
405
pri31057_ch09.pp390-435.qxd
10/15/09
11:02 AM
Page 406
Figure 9.18 The Huaca del Sol at Moche, one of the largest solid adobe structures ever built in the Americas. The back side of the huaca has been washed away by the river, so that only part of it remains today.
_ _
Figure 9.19 Makers’ marks on adobe bricks at Moche.
406
Chapter 9 South America: The Inca and Their Predecessors
millions of adobe bricks, were constructed segmentally in similar, discrete rectangular units composed of stacked bricks. Moseley found that each adobe segment, or column, generally included bricks marked by a single maker’s symbol. Whereas some symbols marked the bricks in numerous construction segments, others appeared less frequently. Thus Moseley suggested that segments of the two pyramids were built by groups of associated laborers, with some work parties involved in the construction more often than others. The use of conscripted laborers, known as the mit’a system, was a familiar means of tribute employed by the Inca more than a millennium later. By A.D. 300–400, Moche influence dominated not only the larger Chicama Valley to the north but also the adjacent Virú and Santa valleys to the south (Figure 9.20). In Virú and Santa, Moche pottery appeared suddenly and rapidly displaced the local pottery traditions. This ceramic shift coincided with dramatic changes in settlement pattern, in which the local populations congregated at a few large settlements. In each valley, a sizable monument was built following the architectural canons used in the two Moche Valley
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 407
A
Figure 9.20 North coastal Peru, with the location of Moche and other valleys dominated by Moche influence.
N
D
E S M
O
PACIFIC OCEAN
U T
A
ue R. m b a yeq Pampa Grande
N
La
IN
S
Archaeological sites Modern cities
R ma . i ca h C
R.
San
M
Galindo e R. Trujillo och . Moche úR V i r R. ta
Ne
0
huacas. Although these pyramids do not approach the Huaca del Sol in size, each was usually the largest structure of its time in its region, and each was associated with Moche pottery or mural art. Farther south in the Nepeña Valley, Moche pottery appeared somewhat later but is limited to settlements in a small area around the site of Pañamarca (pahn-ya-MAR-cah), a large community that may have guarded the southern limits of Moche hegemony. In each of these valleys, Moche influence ceased by A.D. 600, at roughly the same time that the Moche site was abandoned. Before the decline, the Moche people grew cotton, maize, potatoes, peanuts, and peppers by irrigating the coastal desert. Complex canal systems were built to transport water to flat desert lands located kilometers away. The building and maintenance of these water-control systems also may have required labor-recruitment practices similar to those employed in the build-
ñ
a
Pañamarca
pe
150 km
ing of the huacas. Fishing and the hunting of sea mammals provided important dietary resources. To date, large public food-storage complexes, similar to those associated with the later Chimú and Inca polities, have not been found at Moche. Moche society was highly stratified. There were great, and probably inherited, differences in access to wealth and power. Marked disparities have been noted in the quantity and quality of grave goods found within burials. Graves incorporated into or found near the two Moche huacas tended to be more lavish. Status variation also can be inferred from variability in domestic architecture. Residential structures that were more solidly built also were architecturally more elaborate than other domestic constructions; the former tended to be associated with more highly decorated ceramic forms, such as ornate stirrup-spout vessels (Figure 9.21), and other objects of value, such as copper implements.
mit’a system A means of tribute in
prehispanic Andean South America that involved the use of conscripted laborers to complete discrete organizational tasks. stirrup spout A distinctive curving spout on pottery vessels that is shaped like the stirrup of a saddle, characteristic of Moche pottery.
Site Moche
407
_
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 408
Figure 9.21 A Moche stirrup-spout jar with a scene of a ruler receiving a messenger.
Moche artistic expression is amazingly varied. Animals, plants, anthropomorphized demons or deities, and a wide range of life scenes are shown, including the hunting of animals, fishing, combat, the punishment of prisoners and criminals, sexual acts, and the pomp of rulers seated on thrones or carried in litters. Architectural details— temples, pyramids, and houses—are all depicted in Moche art, as are features of clothing and adornment. The degree of realism with which Moche art is expressed and the wide spectrum of subject matter make it one of the most appealing of all pre-Columbian art styles. —Christopher Donnan (1976)
_ _
lost wax casting A technique for casting metal in which a sand or clay casing is formed around a wax sculpture; molten metal is poured into the casing, melting the wax. The cooling metal takes on the shape of the “lost” wax sculpture preserved on the casing.
408
Moche artisans are recognized for their accomplishments in metalworking, making objects of silver and various alloys. They not only used the known techniques of hammering, annealing, embossing, repoussé, and soldering but began to use turquoise mosiac inlay and simple casting as well. They also developed lost wax casting, a technique in which a hollow cast is made by allowing the molten metal to replace a wax sculpture in the cast. Although the bulk of the metal was used for ornaments, copper ingots were smelted for storage. In addition, the Moche smiths made chisels, spear-
Chapter 9 South America: The Inca and Their Predecessors
points, digging-stick tips, spearthrower hooks, tweezers, and fishhooks. The most renowned craftspeople were the potters, who, through a wide array of painted ceramic vessel forms and high relief sculptures, have left an unsurpassed portrayal of Moche mythology, ritual, and daily life. Scenes showing deer being hunted with spearthrowers, as well as fishermen at sea in one-man canoes made of bundles of rushes, are common. Other figures often portrayed include warriors, weavers, prisoners, mothers with children, messengers, and rulers carried in litters or sitting on platforms. Battles are fre-
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
quently depicted, as are scenes in which victorious soldiers accompany naked prisoners with ropes tied around their necks. The pottery also provides further evidence for social stratification. In general, men wore a loincloth and a sleeveless shirt beneath a belted tunic that stopped above the knee. But more important people also wore a large cloth mantle. Although all men wore headgear, only a few important individuals had animal skin headdresses decorated with feathers and small pieces of metal. Moche potters were explicit in their depiction of sexuality, portraying a range of erotic practices. Some of the more graphic depictions are scenes of masturbation and male-female couples engaged in fellatio. Historical data from the time of Spanish contact indicate that the native people on the north coast of Peru practiced sodomy and that some of their ceremonies included orgies. Traditionally, archaeologists and art historians have implied that the Moche ceramicists depicted all aspects of life. Yet, more recently, Christopher Donnan, of the University of California, Los Angeles, has noted that certain scenes, such as the presentation of a goblet to an important person, tend to recur on Moche pottery. This Presentation Theme, which appears to have been part of a Sacrifice Ceremony, usually includes a series of associated symbolic elements (Figure 9.22). At the
Page 409
center of the scene is an important individual with rays emanating from his head and shoulders. The rayed figure generally wears a conical helmet topped by a crescent-shaped ornament and holds the tall goblet (possibly filled with blood). The scene usually also features a warrior who is extracting blood from naked and partially painted captives whose throats were slit. These scenes also generally include a composite human-feline figure who often is holding the captive, animated weapons, and a dog near the rayed figure. Elements of this scene appear individually on other Moche vessels. Recently, Steve Bourget, of the University of Texas at Austin, excavating in the Huaca de la Luna, found the remains of at least 70 males between the ages of 15 and 35. Some of these individuals were intact, and others were dismembered. They appear to represent a number of different sacrificial events. These individuals were found strewn across an earthen-floored patio and confirm that scenes depicted on Moche pottery may correspond to actual events. These findings support Donnan’s earlier proposal that rather than depicting all facets of Moche life, the pottery focuses primarily on a few important ritual themes. This visual record, in conjunction with excavations and settlement pattern studies, provides a more complete picture of the Moche than we have for most other ancient South American peoples.
Figure 9.22 A representation of the Presentation Theme, part of the Moche Sacrifice Ceremony.
_ Site Moche
409
pri31057_ch09.pp390-435.qxd
10/1/09
8:16 AM
Page 410
Concept
<
The Nazca Geoglyphs Desert images
Figure 9.23 An aerial photograph of a monkey at Nazca.
polychrome Multicolored; describing
pottery that has been decorated with three or more colors.
_ _
geoglyph Ground markings, such as the lines and life-form representations found in the Nazca desert.
410
On the south coast of Peru, the Nazca style developed from the earlier Paracas art around 200 B.C. Most Nazca ceramic art, renowned for brilliantly painted polychrome vessels with six or seven colors, comes from looted cemeteries, although a few large Nazca sites have been studied more systematically. Most notable is the site of Cahuachi (cah-WAH-chih) in the Nazca Valley, a large aggregation of adobe platforms, courts, and associated buildings that covers roughly 1 sq km (0.4 sq mi). Cahuachi, with a large 20-m (65-ft)high stepped pyramid, was the largest center of its day in the Nazca region and may have had a role in the dispersal of the Nazca style to other valleys along the south coast. Perhaps the most astonishing feature in the Nazca region of southern Peru is spread out over several hun-
Chapter 9 South America: The Inca and Their Predecessors
dred square kilometers (an area three to four times larger than Manhattan Island) in the desert between the Ingenio and Nazca River valleys. More than 1000 years ago, lines, geometric forms, and various figures were drawn on the desert surface by removing dark rock fragments to expose a light-colored underlying soil. The geometric shapes, mostly triangles and trapezoids, alone cover an amazing 360 ha (900 acres). The majority of the several dozen figures, the most famous of the Nazca geoglyphs, are animals, although several are plants and others are humanlike. The animals include birds, lizards, and fish, as well as a spider and a monkey (Figure 9.23). The latter, an inhabitant of the Amazon rain forest and not of the dry coast, is known to have been important in coastal religion and symbolism. One of the myths about the Nazca geoglyphs is that they must be seen from the air to be appreciated. More recent studies indicate that they were meant to be viewed by people on the ground. Most of the lines and drawings are visible from the desert floor, and they can be seen even more clearly from either the low hills or the huge sand dunes that border the dry plain. Although the specific intentions of the original makers remain a matter of some debate, recent systematic research of the geoglyphs by Markus Reindel, of the German Archaeological Institute, and colleagues has revealed that the desert features were walked by the Nazca. As these lines and features were walked, the Nazca left small offerings in selected locations. Reindel suspects that the geoglyphs were part of a ritual landscape associated with water and fertility rites.
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 411
Site
:
Sipán Clues to Moche ritual and politics
Less than 200 km (124 mi) north of the pyramids of Moche is the fertile Lambayeque (lahm-bye-YEA-kay) Valley. Abundant water resources and rich, level farmland have combined to make this area highly suitable for intensive agriculture. At the center of the valley lies the small village of Sipán (see-PAHN), an agricultural community of approximately 1500 people, distinguished by the presence of ancient pyramids and cemeteries. The impressive Sipán pyramids are visible for miles, rising above the intensively farmed fields that today grow mostly sugarcane (Figure 9.24). For decades, some of the farmers of Sipán have supplemented their agricultural income with sporadic looting from the rich archaeological resources in the region. The looters dig deep holes with picks and shovels and often find ceramic vessels or stone and shell beads, which they sell to the antiquities dealers who periodically pass through the region. Occasionally, the looters expose tombs containing objects of gold or silver, for which they receive much larger sums. Such occasional rich finds, and the periodically weak economy in the Lambayeque area, have helped keep this destructive tradition of looting alive. Over generations, some Sipán inhabitants have become adept at learning where to dig to recover the most valuable items and at avoiding the local police and government officials responsible for protecting archaeological resources. At the end of 1986, the local looters, believing they had exhausted nearby cemeteries, focused their attention on the smallest of three Sipán pyramids. Working at night to avoid detection, they dug deep pits into the pyramid’s core of solid mud bricks. For months, little was found. But on the night of February 16, 1987, they tunneled into
one of the richest burial chambers ever looted, the royal tomb of a Moche ruler. Conflicting accounts make it difficult to know exactly what was removed, but several cloth sacks of priceless artifacts were undoubtedly carted away. Almost immediately, disputes arose over the distribution of the treasure, and one looter, displeased with his share, informed the police. The police went to Sipán, where they arrested some of the other participants and confiscated part of the plunder in their homes. They also notified Walter Alva, who had worked
Figure 9.24 An aerial view of the three pyramids at Sipán. The smallest pyramid, in the lower left, was the focus of looting and subsequent systematic excavation.
_ Site Sipán
411
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 412
Figure 9.25 Excavations directed by Walter Alva in Tomb 1 in the summit of the smallest pyramid at Sipán.
_ _
Figure 9.26 The six building episodes of the small pyramid at Sipán. The pyramid, beginning as a low platform, was enlarged during each episode of construction. The Old Lord of Sipán was buried in a tomb in the lowest level.
412
Chapter 9 South America: The Inca and Their Predecessors
for years with the police in an effort to halt the looting. Alva recognized the richness and scholarly significance of the finds, as well as the need to locate the tomb. In Sipán, word spread quickly about the discovered wealth, and many local residents flocked to the pyramids with window screens, kitchen utensils, and shovels to see what they could recover. Even after the police reached the scene, it took several hours to clear the pyramid of wealth seekers. From that point to the conclusion of the scientific excavations, the exposed pyramid had to be guarded 24 hours a day. The damage to the funerary chamber was severe, from both the original grave robbers and those who later streamed to the site and ruthlessly hacked at the walls. Alva decided that an archaeological project should start immediately to investigate the pyramid before this important site was entirely destroyed. Although some of the looted finds were recovered that first night and a few items were confiscated later by the police, most of the plundered objects were never found. Instead, they were surreptitiously smuggled to Lima. From there, many items eventually were illegally transported to the
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
art markets and private collections of Europe, the United States, and Japan. Walter Alva, Christopher Donnan, and their colleagues immediately set to work on the Sipán pyramids (Figure 9.25). They mapped the complex, thereby determining the relationship between the smallest structure (now pitted) and the larger pyramids. Because the small mound was in greatest danger of further destruction, they concentrated their efforts there. Detailed architectural analysis of the areas exposed by the looters’ trenches led to the discovery that the pyramid was built in six stages (Figure 9.26). In its earliest phase, thought to have been constructed during the first century A.D., the pyramid appears to have been only a low rectangular platform with two steps extending along its entire north side. Each subsequent building enlarged the pyramid, encapsulating the previous structure, with the final
Page 413
construction phase completed around A.D. 300. The richness of the looted finds, as well as elaborate gold, silver, and ceramic artifacts recovered during the cleaning operation, prepared Alva and Donnan for the possibility that an important burial might be recovered. Before Sipán, the Moche burials with the most elaborate pieces generally had been looted rather than excavated scientifically. Yet the excavators could never have imagined that they would recover perhaps the richest burial ever found in the Western Hemisphere. Nor could they have known that their findings would help answer a key riddle about the ancient Moche. Were the repetitive scenes recognized in Moche art, such as the Sacrifice Ceremony (see p. 409), really enacted by the Moche people or purely mythic in nature? Over months of painstaking effort, the Sipán excavations uncovered three
Figure 9.27 A large, gold, crescentshaped headdress, 62.7 cm (24.7 in) wide, from Tomb 1 at Sipán.
_ Site Sipán
413
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Figure 9.28 A miniature figure from the tomb of the Old Lord at Sipán, holding a war club (5 cm, or 2 in, tall) and a shield. He wears a headdress of thinly hammered gold with an owl in the center and a gold nose ornament. Turquoise adorns his body, ears, and eyes.
y
The looting of archaeological sites is a major problem in many areas of the world. What do we lose in addition to the objects themselves?
scepter A staff or baton borne by a
_ _
ruler as an emblem of his/her position and authority.
414
Page 414
fabulous tombs. In each case, the central figure was elaborately costumed in exquisitely crafted gold and silver ornaments, as well as worked shell, gemstone, and metal finery of all kinds. In the two largest, most elaborate, and latest tombs—Tombs 1 and 2—the central figures had key costume elements that enabled the excavators to identify them as specific participants in the Sacrifice Ceremony. For example, the central rayed figure, referred to as the Warrior Priest, in the Presentation Theme often was depicted with paired backflaps, a crescent-shaped headdress piece, large circular ear decorations, a crescent nose ornament, and a pair of bells hanging from his belt (the large figure on the left in Figure 9.22). This individual is usually shown in military settings, accompanied by a dog, and often with a lamp-shaped scepter or war club. All these elements were excavated together in Tomb 1 at Sipán (Figure 9.27). Tomb 2 yielded an
Chapter 9 South America: The Inca and Their Predecessors
individual buried with many of the decorative and symbolic elements that have been linked to a second principal figure in the Sacrifice Ceremony, the Bird Priest (to the right of the Warrior Priest in Figure 9.22). The earliest of the three excavated tombs (Tomb 3) was neither as large nor as closely associated with the depictions of the Sacrifice Ceremony as the later two. Yet the central figure, called the Old Lord of Sipán by the site’s excavators, was richly entombed with some elements of the Warrior Priest attire (Figure 9.28). It is possible that the Sacrifice Ceremony was less elaborate and stylized in earlier Moche times; artistic depictions of this event all postdate A.D. 300. Amputated hands and feet were uncovered at Sipán. These sacrificial offerings would seem to indicate, not only that key participants in the Sacrifice Ceremony were buried in the pyramid, but also that the event itself was probably practiced on or near the pyramid. The Sipán excavations provide a clearer understanding of Moche social and economic organization. Moche society was marked by greater differentials in wealth than previously believed. The elaborate costumes buried with these nobles would have demanded many highly skilled artisans to replenish the great objects of wealth that were buried. This burial custom helps contextualize the extraordinary artistic and technological achievements long associated with Moche craftworkers. The royal tombs also have yielded significant clues about Moche art and religion. We now recognize that at least some of the art documents actual events enacted by real people. Although the treasures of the Sipán tombs seem priceless, of greater value is the exciting new cultural information they have provided, enabling us to develop a clearer picture of the Moche—one of the most fascinating civilizations of the ancient world.
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 415
Site
:
Tiwanaku Bolivia’s high-altitude ancient city
When the Spanish chroniclers asked the Inca rulers about their origins, they were told that the genesis of the first Inca took place in the part of the realm called Collasuyu (coy-yah-SUE-you). This Inca origin myth may have developed from the belief in the sacred nature of Collasuyu’s Lake Titicaca, the largest lake in South America. When
the great king Pachakuti (A.D. 1438– 1470) began the Inca imperial conquests, it is no coincidence that he moved first against polities on the shores of Lake Titicaca. The first sedentary settlements south of Lake Titicaca, in what is today Bolivia, were established during the second millennium B.C. The inhabitants of
50 km
0
N
[E]ven the great Emperor Inca Pachacuti was impressed by the stonework at the ruins of Tiwanaku in the late 15th century . . . the emperor ordered his engineers to learn the stone masonry techniques and use them in Cuzco. This account is most likely apocryphal but we can consider it the first recorded instance of the site being used for political purposes. Since at least this visit to the site by a conquering Inca emperor to the present day, Tiwanaku has stood at the center of some highly charged political, ideological, ethnic, and scientific debates. —Charles Stanish (2002)
ECUADOR Lake Titicaca
PERU
Chiripa
BRAZIL
Pampa Koani Tiwanaku
Wari
Wari site
Lake Titicaca
Tiwanaku site
Tiwanaku
Wari empire A
PACIFIC OCEAN
BOLIVIA
no la ip lt
Tiwanaku empire
T r o p i c o f C apr ic o r n
CHILE ARGENTINA 0
300 km
Figure 9.29 Lake Titicaca and the domains of Tiwanaku and Wari. Site Tiwanaku
415
_
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 416
Ancient lake Di tch
North gate Kalasasaya N
Sunken temple
Akapana 0
Dit
150 m
ch
Pumapunku
Figure 9.30 The plan of the civicceremonial core of Tiwanaku.
_ _
altiplano (Spanish) The high-altitude plain between the eastern and western ridges of the Andes in Peru.
416
these early communities on the windswept high-altitude steppe, or altiplano, had a mixed agrarian economy in this zone that most North Americans would consider to be too elevated for agriculture. In the cold altiplano, the early villagers subsisted on a range of domesticated species adapted to the torrential rains of the wet season (November– May) as well as an extended dry season (April–October). The most important food resources were hardy tuberous plants like the potato, cold-adapted grains like quinoa, and such domesticated camelids as alpaca and llama. One of the early villages on the southern shores of Lake Titicaca was Chiripa (cheer-EE-pah) (Figure 9.29). At the center of the site was a large artificial mound that was constructed around 900 B.C., largely contemporaneous with Chavín de Huántar in the Andean highlands to the north (although the site itself was first occupied centuries earlier). A series of subterranean houses were arranged around the mound. During recent excavations at the site, Christine Hastorf, of the University of California at Berkeley, and her colleagues discovered a semisubterranean structure (with a plaster floor), 200 meters south of the
Chapter 9 South America: The Inca and Their Predecessors
main mound, that dates to 800–750 B.C. This court, which measures approximately 11 ⫻ 13 m (36 ⫻ 43 ft), is one of the earliest known semisubterranean features in the region and presages the much larger sunken courts that occur later at Tiwanaku. Around 800 B.C., a small settlement was founded 20 km (12 mi) south of Lake Titicaca, on the site that was to become the great city of Tiwanaku (teewah-NOK-u). Situated 3850 m (12,600 ft) above sea level in a small valley in the Bolivian altiplano, at more than twice the altitude of Denver, Colorado, Tiwanaku was one of the highest urban centers ever established. Monumental architectural construction began during the first centuries A.D., when the site rose to importance in the southern basin of Lake Titicaca. Tiwanaku grew to its greatest geographic size (more than 4 sq km, or 1.5 sq mi) and demographic extent (an estimated 25,000–40,000 people) around A.D. 400. For the next 600–800 years, the city dominated the entire Lake Titicaca region, and ceremonial art styles associated with the capital spread over a far wider area. The civic-ceremonial core of Tiwanaku was a 20-ha (50-acre) area at the center of the site. This precinct, laid out according to a grid oriented to the four cardinal directions, included a series of truly impressive architectural features (Figure 9.30). Akapana (ah-kah-PAHnah), an enormous stone-faced, stepped platform mound, 200 m (655 ft) on a side and 15 m (50 ft) high, was the most massive construction. It included huge blocks of stone, some of which weighed as much as 10,000 kg (11 tons) and had to be brought over land and water to Tiwanaku from quarries more than 100 km (62 mi) away. A second platform, Pumapunku (pu-mah-POON-ku), measured more than 5 m (16 ft) high, with walls 150 m (500 ft) long. Tiwanaku’s civicceremonial core also included a complex of buildings, with a stone drainage system, that may have been a palace. The most famous stone sculpture at Tiwanaku, the Gateway of the Sun, was incorporated into a large rectangular raised platform known as the Kalasasaya (kah-lah-sah-SIGH-yah).
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 417
Figure 9.31 The Gateway of the Sun at Tiwanaku, shown here, broken in two, before it was excavated and the two halves were realigned. The central figure above the gate holds two scepters that end in the heads of condors.
The Gateway of the Sun, carved from a single huge stone block, portrays a central figure holding two scepters that end in the heads of condors (Figure 9.31). This individual is flanked by rows of winged attendants who each carry a condor scepter. The central character recalls a figure with two staffs, frequently shown in Chavín iconography. Inside the Kalasasaya are rectangular sunken courts (Figure 9.32). These architectural features, which may have been temple precincts, have a precedent at earlier Chiripa. At Tiwanaku, the local altiplano economy of camelid pastoralism and the cultivation of hardy grains and tubers was supplemented by the reclamation for agriculture of waterlogged lands adjacent to Lake Titicaca. In addition, Tiwanaku maintained longdistance trade connections, establishing economic colonies in both the Pacific Coast zone to the west and the more tropical forested zones to the east. These earlier networks may have laid a foundation for the Inca to link the diverse and distant reaches of the Andean realm. Warm-region crops, such as maize and coca, were obtained, as were birds and medicinal herbs from the east and obsidian and coastal products, such as shell and dried fish, from the west. Trade connections were maintained by
large llama caravans, which traveled throughout the Bolivian altiplano and to southern Peru and the Chilean coast. Alan Kolata, of the University of Chicago, has studied agrarian reclamation on the Pampa Koani (pahmpah ko-AH-nee), an area subject to frequent natural flooding on the southern border of Lake Titicaca. During the first millennium A.D., the waterlogged land was reclaimed through the construction of raised fields near the lake, by excavating the heavy Pampa soils on either side of the field surface and mounding the dirt in the center to form an earthen platform elevated above the seasonally high waters. Kolata found that some raised platforms were constructed more elaborately to enhance plant growth. In one field at the lakeshore, the platform consisted of a cobblestone base, covered by a thick clay stratum, superseded by three gravel layers and a rich layer of topsoil. In this ingenious structure, the clay, held in place by the cobblestones, prevented the brackish lake water from penetrating into the field from below, while fresh rainwater could percolate down to the roots of the crops from above. During the latter half of the first millennium A.D., while Tiwanaku was the principal center in the southern Andes,
On January 23, 2006, Evo Morales was formally inaugurated as Bolivia’s new president in the capital city of La Paz. . . . But for many Bolivians, his “spiritual” inauguration the day before at the archaeological site of Tiwanaku represented the watershed moment in recent Bolivian political life. Evo Morales is the first indigenous president of Bolivia, and the first indigenous national leader in South America since the defeat of the Inca Empire by a small Spanish army almost 500 years ago. —David Kojan (2008)
coca A native Andean shrub whose
dried leaves are chewed as stimulants.
Site Tiwanaku
417
_
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Figure 9.32 Cut-stone walls that line the sunken court of the Kalasasaya at Tiwanaku. The carved stelae and the anthropomorphic figures incorporated into the walls are part of Tiwanaku’s powerful iconography.
_ _
Figure 9.33 A pottery vessel decorated with Tiwanaku/Wari stylistic elements.
418
Page 418
the Peruvian highland site of Wari (WAR-ee) dominated the more northerly highland and coastal regions (see Figure 9.29). Andean archaeologists suggest that whereas Tiwanaku developed economic ties with its hinterland, Wari’s control over the surrounding area was more militaristic. Wari’s regional domination also was shorter, since that site lost its influence and was largely abandoned before the tenth century A.D. Although the artistic styles associated with Tiwanaku and Wari were similar, both of them featuring jaguars and raptors (birds of prey) (Figure 9.33), the specific relationship between the two centers remains unknown. Soon after A.D. 1000, Tiwanaku’s domination over the basin of Lake Titicaca began to wane. A number of smaller competing states emerged, each maintaining its own local sphere of influence until the middle of the fifteenth century A.D., when the Titicaca Basin was unified under Inca domination. The decline of Tiwanaku coincided with the abandonment of the raised fields of the Pampa Koani. The vast landscape of once-productive fields reverted to pasturage, a function they still serve today. Even the Inca
Chapter 9 South America: The Inca and Their Predecessors
did not reclaim the seasonally waterlogged Pampa lands that had once helped feed the population of Tiwanaku, focusing instead on the rocky mountain slopes above the plain. There, the Inca expanded the region’s agricultural productivity by constructing terraces that transformed steep hillsides into arable farmland. Although the specific causes of Tiwanaku’s collapse have not been determined, we do know that several of the important organizational features associated with the later Inca empire can be traced to this earlier altiplano center. The widespread Tiwanaku iconographic style seems to presage the broad dissemination of Inca state art. Likewise, Tiwanaku’s establishment of small economic colonies in diverse environmental settings augurs an Inca practice. Perhaps the Inca kings recognized the powerful influence of Tiwanaku on the later prehispanic Andean world when they traced their ancestry to the basin of Lake Titicaca. In so doing, the Inca rulers also were invoking the mystique of Tiwanaku’s past grandeur to bolster their own efforts to reestablish a farreaching Andean domain.
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 419
Site
:
Chan Chan Desert city of the Chimú
ECUADOR
PERU yeque mba La
C
R.
R ma . ca i h
Chan Chan PACIFIC OCEAN
V
R. irú
Moche R.
nta R.
Sa
On the north coast of Peru, the seventhcentury collapse of Moche coincided with a shift in political power. In the Moche Valley, occupation concentrated in the large settlement of Galindo (gahLEAN-doh). Yet unlike at the Moche site with its huge adobe huacas, the nondomestic construction at Galindo was small. At the same time, a much larger settlement, Pampa Grande (pahm-pah GRAHN-day), with more elaborate architecture, was established in the Lambayeque Valley to the north (see Figure 9.20). Both Pampa Grande and Galindo were positioned in easily defendable locations from which they could control local irrigation systems. Thus, although the north coast was not conquered by Wari invaders from the south, that expansionist polity may have caused sufficient unrest to effect major shifts in population and power. By approximately A.D. 800–900, a small settlement had been founded at the mouth of the Moche Valley at Chan Chan (CHAHN chahn). Although the exact extent of this early occupation has not been determined, it is clear the site grew rapidly in size and importance. By the middle of the fifteenth century, Chan Chan had become a sprawling coastal city, covering more than 20 sq km (8 sq mi). Chan Chan was also the capital of the Chimú (chee-MOO) state, which stretched 1000 km (620 mi) from southernmost Ecuador to central Peru (Figure 9.34). At its height in the mid-fifteenth century A.D., the civic-ceremonial core of Chan Chan covered roughly 6 sq km (2.3 sq mi, roughly twice the size of New York’s Central Park). This central area was dominated by 10 rectilinear compounds, or ciudadelas (literally, “little cities”), each surrounded by high adobe walls (Figure 9.35). The nucleus of Chan Chan also included a large platform-court complex, as well as flat-
Nepeña R.
Extent of Chimú empire Chimú or Chimú-occupied site Modern international border Chicama-Moche canal
0
150 km
topped mounds and numerous smaller monuments and structures; yet the most striking and massive features were the ciudadelas, many of which measured 200–600 m (650–1950 ft) on a side. The ciudadelas, oriented roughly north-south, each had a single entrance through the north wall that opened onto a corridor leading to a broad court (Figure 9.36). A ramp
Figure 9.34 The Chimú empire stretched along the Pacific Coast of South America from southern Ecuador to central Peru.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on Chan Chan, see the Internet exercises on your online learning center.
Site Chan Chan
419
_
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 420
N
0
80 m
PACIFIC OCEAN
Figure 9.35 The plan of central Chan Chan, which is dominated by the ciudadelas. The Rivero ciudadela is in the bottom center of the plan.
_ _
split inheritance An Andean practice by which the successor to the throne inherited only the office of the dead ruler; his junior kinsmen received the lands, palace, and personal wealth of the dead ruler.
420
sloped up to a long bench along the south wall of the court. Human corpses, probably buried when the compounds were constructed, were uncovered during excavations of these ramps. Each compound also included a multitude of storerooms, smaller courts, and living quarters (Figure 9.37). Usually the last structure built in each ciudadela was a burial platform, which presumably held the bodies of the family that occupied each of these mazelike compounds. Later records name ten Chimú kings, corresponding to the ten palatial compounds at Chan Chan. Geoffrey Conrad, of Indiana
Chapter 9 South America: The Inca and Their Predecessors
University, believes that each ciudadela was a particular king’s residence and the administrative center for the Chimú kingdom. Following the death of the ruler, the compound became the mausoleum for the king, maintained by his living kinsmen. Conrad argues that the maintenance of a dead Chimú ruler’s personal property by his heirs, who did not succeed him in office, was derived from a broader pan-Andean custom of ancestor worship. Upon succession, each new ruler built his own ciudadela to serve as headquarters and royal treasury. Most of the inhabitants of Chan Chan resided outside these massive compounds in small structures composed of six to eight small rooms. During excavations in these quarters, few farming or fishing implements were found. Instead, the evidence for craft manufacture was abundant, including lapidary and woodworking tools, spinning and weaving implements, and the equipment for metalworking. As at earlier Moche, most of the monumental construction at Chan Chan was built by a labor force from outside the center. Michael Moseley estimates that the population at Chan Chan remained relatively small, perhaps 25,000 people. In the city’s hinterland, huge labor investments were made in agricultural intensification. Sophisticated hydraulic systems brought water to the land surrounding Chan Chan. The Chimú even constructed an intervalley canal designed to carry water to the Moche Valley from the Chicama River, 65 km (40 mi) away. The development of the large irrigation networks around Chan Chan would seem to relate to the Andean practice of labor tribute (the mit’a system), as well as to the Chimú pattern of succession. According to one hypothesis, the Chimú, like the Inca, practiced a policy of split inheritance, by which the successor to the throne received the inherited office of the supreme leader, but the land, the palace, and the personal wealth of the dead ruler were left to a corporate group of other junior kinsmen. As a consequence, each new ruler had to raise his
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 421
Figure 9.36 The plan of the Rivero ciudadela at Chan Chan. The U-shaped structures, audiencias, may have been used by nobles holding audience with lower-status people. Each structure holds only one seated person, and each has its own court.
Annex
How did the built environment at Chan Chan compare with that of earlier Moche? What might these differences mean in terms of past behaviors?
y
Burial platform N
Audencia Storerooms
0
100 m
own revenues to erect his residential compound and finance his reign. His principal resource was a large labor force that could be employed in monumental construction, agricultural intensification, road building, and militaristic ventures. Between 1462 and 1470, the Chimú were in competition with the increasingly powerful Inca. By the end of the decade, this conflict had ended with the
Walk-in well
incorporation of the Chimú kingdom into the rapidly growing Inca empire. In this manner, the Inca were able to link their lands and road systems with those previously controlled by the Chimú. Curiously, Chimú artifacts were found more widely distributed after the Inca conquest than before, perhaps as a result of the emergence of new patterns of trade and taxation, or possibly because of the great admiration the Inca had for
_ Site Chan Chan
421
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 422
Figure 9.37 This U-shaped structure, or audiencia, at Chan Chan, with two large niches in each interior wall. Elaborate decorative carvings adorn the interior adobe walls of the audiencia.
Figure 9.38 An aerial view of the western half of central Chan Chan. The large compound in the right foreground is the Rivero ciudadela.
Chimú craftspeople. For example, Chimú metalworkers were brought to the capital, Cuzco, to work for the Inca state. Today the quiet desolation of the dry coastal environment is broken only by the nearly continuous howl of onshore winds. No one lives at Chan Chan, but the site is covered by a series
_ _
422
Chapter 9 South America: The Inca and Their Predecessors
of trails and paths that skirt the undulating tops of collapsed adobe ruins that were once the palaces of the Chimú kings (Figure 9.38). At dawn from a distance, the diffused, weak daylight makes it difficult even to distinguish the massive earthen walls of this once important city from the sky at the horizon.
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 423
Site
:
Cuzco and Machu Picchu Highland Inca settlements drive them far from the Inca homeland. Cusi Inca Yupanqui was crowned Inca and renamed Pachakuti, “he who remakes the world.” From that date, generally placed around A.D. 1440, Pachakuti is said to have initiated the dramatic series of conquests that culminated less than a century later in a huge empire that stretched over 4200 km (2600 mi) from north to south. Inca imperial propaganda proclaimed that the Andean world was in a state of savagery before the rise of the Inca. This claim was obviously ethnocentric and fallacious. Clearly, the royal
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on a research expedition to find Inca mummies, see the Internet exercises on your online learning center.
Figure 9.39 The Inca empire (Tawantinsuyu) was divided into four quarters: Chinchaysuyu in the north, Antisuyu and Cuntisuyu in the center, and Collasuyu in the south.
CHINCHAYSUYU Lambayeque R. Chicama R. Moche R. Huánuco Pampa Nepeña R. ChillónR.
Machu Picchu
ANTISUYU Cuzco
Nazca R.
CUNTISUYU
Lake Titicaca
in Bas ca ica o lan tip Al
Tit
As with the Aztecs of Mesoamerica, much of what we know about the Inca comes to us not from archaeology but through written documents, filtered through the eyes of European chroniclers, whose goals and experiences differed greatly from those of the Inca themselves. In 1532, a small band of Spaniards led by Francisco Pizarro came into contact with populations in the northern coastal valleys of Peru that were part of a giant centralized political domain called Tawantinsuyu (Figure 9.39). The capital of this great polity was Cuzco (KUSE-co), a city in the southern highlands. The highestranked leader was called Inca, son of the deity Inti (the Sun God) and descendant of a long line of heroic dynasties. At that time, the empire was in the midst of a severe crisis, provoked by a bitter rivalry between two brothers, Huascar and Atauhualpa. Although Atauhualpa eventually won the power struggle, his costly victory may have lost the empire; control of the Inca domain was soon in the hands of the Spanish. According to legend, the rapid rise of the Inca began with hostilities between the inhabitants of Cuzco and the Chanca, a neighboring people. Fresh from military victories over several adjacent polities, the Chanca laid siege to Cuzco, forcing many of the inhabitants, including the reigning Inca, to flee to the surrounding hills. At Cuzco, a son of the Inca ruler, named Cusi Inca Yupanqui, was left to spearhead the final defense of the city. While waiting for the last Chanca onslaught, Cusi Inca Yupanqui had a vision in which he was told by a supernatural being that he would eventually rise to power and conquer many nations. Inspired by the apparition and buoyed by the support of new allies, Cuzco’s defenders rallied to defeat the Chanca and
PACIFIC OCEAN
Tropic of Ca
0
pricorn
COLLASUYU
250 Miles
_ Site Cuzco and Machu Picchu
423
pri31057_ch09.pp390-435.qxd
10/1/09
Figure 9.40 The north face of the fortress of Sacsahuaman at Inca Cuzco, Peru. This edifice was constructed of huge stone blocks fitted together with great masonry skill. Some of these blocks weigh more than 100 metric tons. The fortress is perched atop a high hill that has a commanding view of Cuzco below.
_ _
frieze A decorative band or feature, commonly ornamented with sculpture, usually near the top of a wall.
424
8:17 AM
Page 424
families at Cuzco inherited administrative strategies and systems of legitimization from the rulers of earlier states, centered at Chan Chan and Tiwanaku. Archaeology has shown that at the outset of the second millennium A . D ., following the fall of Tiwanaku, the southern highlands around Lake Titicaca did undergo an episode of decentralization in which many small polities fragmented a landscape that had previously been unified by the earlier altiplano center. In support of the legendary accounts, archaeological analyses indicate that by A.D. 1450, these autonomous competing polities of the Titicaca Basin had begun to come under Inca hegemony. Centered at Cuzco, the Inca rapidly established major ceremonial centers on islands in Lake Titicaca, not only to solidify their control of this important geographic resource, but also to legitimize their ancestral link to earlier Tiwanaku. According to tradition, Cuzco was established by Manco Capac, the first Inca ruler, in a mountain valley nearly 3500 m (11,500 ft) above sea level. Yet earlier excavations in conjunction with recent archaeological surveys have established that the site was one of the largest in the region as early as A.D.
Chapter 9 South America: The Inca and Their Predecessors
1000. Because Cuzco is relatively close to the equator, its climate is reasonably mild in spite of the high elevation. Cuzco remained a small community until it was rebuilt by the victorious Pachakuti. The Inca canalized and straightened several small rivers that cut through the settlement. Cuzco’s most elaborate buildings, many of which were constructed entirely of finely hewn stone, were erected in the area between these rivers. An imposing fortress with massive masonry walls, the Sacsahuaman, was built on a steep hill above this central area, which was not fortified (Figure 9.40). Drawing on Inca sources, John Rowe, of the University of California, Berkeley, has suggested that Cuzco was intentionally laid out in the shape of a puma, with the fortress representing the animal’s head and the intersection of the rivers serving as its tail (Figure 9.41). The area between the puma’s front and back legs was paved with pebbles and served as a central ceremonial square. The magnificent Temple of the Sun, the most important Inca structure in Cuzco, was built by Pachakuti at the site’s center. The building’s exterior walls, more than 50 m (160 ft) long per side, were decorated with a thick gold frieze. Entrances to the structure were covered with heavy gold plates, and the interior walls also were coated with gold. Early Spanish accounts describe a central patio or garden that included both natural and gold-crafted vegetation, including maize plants in which the ears, stalks, and leaves were all made of metal. The temple’s principal sanctuary was dedicated to Inti. Inside the room were idols of gold and silver, as well as mummies of former Inca rulers and their wives. Cuzco also contained other, smaller temples, in addition to public buildings and elite residences. Generally, these structures were built of cut stone, finished with such precision that the joints matched perfectly and mortar was not required. These architectural monuments, which required both great skill and significant labor power, were
u
Sacsahuaman
h in
C
erected by commoners as part of the mit’a system. Pachakuti also erected a special public building that served as a convent for Chosen Women, brought to the capital to weave the finest cloth and to participate in specific religious rituals. As the nucleus of an empire, Cuzco was not a typical Inca city. Its principal function was administration, and many of its inhabitants were involved in civic-ceremonial activities: priests, nobles, military officers, architects, and servants. Most of the supporting commoner population resided in smaller communities in the surrounding valley. In these settlements, the structures were built of fieldstone and adobe rather than cut-stone blocks. A network of state storehouses and religious shrines also marked the well-planned terrain in the immediate vicinity of Cuzco. The market in Cuzco was small and outside the center of the city. The peripheral nature of commercial activities at Cuzco reflects the administrativeelite character of the capital. However, it also indicates the nature of Inca economy, in which the state had a central role in the collection and redistribution of goods. Craftworkers and other specialists contributed labor to the imperial government, which thus took control of a wide range of craft items. Compared with the contemporary Aztec polity, private trade and marketing occurred at very low levels in the empire of the Inca. The Spanish were greatly impressed by Cuzco, its stone architecture, and its treasure of precious metals. Before the arrival of the Europeans in 1532, the accumulation of wealth was fostered by an Inca regulation that forbade anyone to remove gold, silver, or fine cloth from the city, once it was inside. In 1535, Cuzco was systematically burned and largely destroyed by the Inca themselves, when they besieged a Spanish garrison. But because the conquering Spaniards reconstructed the city promptly, often using Inca builders, much of the prehispanic plan has endured. Solid Inca walls still provide the
Page 425
uy
8:17 AM
tis
10/1/09
An
pri31057_ch09.pp390-435.qxd
ch
ays uyu
Collasuyu
Temple of the Sun u
uy
is nt
Cu
stone foundation for many later buildings. Even the remnants of the oncespectacular Temple of the Sun now stand as the base for the Church of Santo Domingo. Like modern Mexico City, lying above the ruins of ancient Aztec Tenochtitlán, Cuzco is undoubtedly one of the longest continuously occupied cities in the New World. Today, the most famous archaeological site in Peru, and perhaps all of South America, is another Inca site, Machu Picchu (MA-chew PEE-chew). This site is dramatically situated on the eastern frontier of the Inca empire, and the ruins lie on a saddle between two mountain peaks, which tower over the site and overlook a large bend in the Urubamba River (Figure 9.42). Machu Picchu is a graphic reminder of the architectural grandeur of the Inca. Yet, in spite of its international popularity as
Figure 9.41 The plan of Colonial period Cuzco, which was built on Inca remains. The fortress and walls of the Inca city are outlined as a stylized puma.
What contrasts can you draw between the political structures of the Aztecs and the Inca?
y
_ Site Cuzco and Machu Picchu
425
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Figure 9.42 The superbly preserved royal estate of Machu Picchu, high in the Peruvian Andes. Made known to the outside world early in the twentieth century by Hiram Bingham, the Machu Picchu ruins are a conglomerate of courtyards and stone stairways, as well as residential structures and public buildings.
_ _
archaeoastronomy The study of ancient alignments and other aspects of the archaeological record and their relationship to ancient astronomical knowledge and events.
426
Page 426
a tourist attraction, archaeologists still have much to learn about this magnificent ruin. Burials at the site included both males and females, suggesting that, despite its remote location, Machu Picchu was not simply a military garrison or fortress. Nevertheless, access to the site was limited by gates and removable rope bridges. The ceremonial core of Machu Picchu could be entered only through a stone gateway, making access to this part of the site especially difficult. According to Spanish legal documents from the time of conquest, Machu Picchu was the property of the descendants of the Inca ruler Pachakuti. Machu Picchu was one of several royal estates built by Pachakuti as commemorative memorials to his various military/political exploits, and it was passed to his kin at his death. Elaborate dwellings at the site housed the king and his court when they were
Chapter 9 South America: The Inca and Their Predecessors
in residence, usually during the dry season. A smaller permanent population, responsible for maintaining the estate, may have resided at the site during the rest of the year. The site’s large number of sacred features, as well as its location, reflects Pachakuti’s concern with religion. The topography of the area—an impressive canyon surrounded by large rocks, conical peaks, and snowy vistas—combines many features that were important to Inca religion and cosmology. At the site, an intact intihuatana, a large stone pillar thought to have had a ceremonial function, was found. Usually, when the conquering Spanish found such features at Inca sites, they destroyed them because of their suspected association with the native religion. Some researchers have proposed that the intihuatana at Machu Picchu was used by Inca priests as a sundial to read the length of the shadows cast by the sun. However, researchers specializing in archaeoastronomy have found that the pillar has no obvious astronomical function. Nevertheless, other features at Machu Picchu did have astronomical functions. A cave just below the site is thought to have been an observatory for the December solstice, and from the Torreón, a large masonry tower built on a rock promontory with a clear view to the east, the Inca could record the path of the rising sun during the June solstice. Although Machu Picchu was inhabited by the Inca late in the prehispanic era, it may not have been visited by the Spanish conquistadores. In fact, unlike Cuzco, Machu Picchu was largely unknown to the Euro-American world until it was publicized following an expedition by Hiram Bingham in 1911. Today, amid the clouds and the snowcapped mountain peaks, Machu Picchu remains a magnificent remnant of a once great and powerful civilization that ruled the prehispanic Andean world.
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 427
Concept
<
Inca Highways Lifeline of the Inca empire
The movement of goods and information between the Pacific Coast, the Andean highlands, and the Amazonian montaña has been an important aspect of the South American world since at least the first millennium B.C. John Murra has shown that many Andean ethnic groups established colonies and devised other strategies to gain access to the products from these diverse ecological niches. As in Mesoamerica, much of this long-distance traffic was accomplished by foot. Yet in prehispanic South America, light loads (up to 45 kg, or 100 lb) also were carried by the domesticated llama, slightly easing the cost of long-distance transport. Archaeologists do not know when the first Andean roads were built. Some scholars have suggested that they date to the time of Chavín de Huántar. By the time Tiwanaku and Wari dominated the Andean landscape, roads were in use. Well-built roads cross right through major Wari centers. Later the Inca linked existing roads into their own extensive system, creating South America’s largest contiguous archaeological remain (Figure 9.43). Not surprisingly, Cuzco was the hub of the Inca road system. The principal road ran through the highlands from the capital to Quito (in Ecuador) and into present-day Colombia. A second mountain segment ran south through Bolivia to Santiago (Chile) and northwest Argentina. A long coastal road connected with the highland network at more than a dozen points. The total road system spanned more than 40,000 km (25,000 mi). Wheeled transport did not exist, and the Inca roads generally were narrow, ranging from 1 to 16 m (3 to 52 ft) wide (Figure 9.44). For the most part, the mountain roadways were more winding and narrower than those along the coast, to accommodate the
terrain. In some places, the road was little more than a well-beaten path. Deserts and swamps were avoided when possible, although the Inca had means for coping with all these situations. Across the desert, roads were lined with walls of posts, adobes, or stones to block windblown sands. In
Figure 9.43 The Inca road system stretched along the Pacific Coast of South America from Ecuador into Chile and northwestern Argentina.
COLOMBIA Quito ECUADOR
PERU BRAZIL
Huánuco Pampa Machu Picchu Cuzco Lake Titicaca BOLIVIA
CHILE Tropic of C
apricorn
PACIFIC OCEAN
Archaeological sites ARGENTINA
Modern cities Major roads
Santiago
Minor roads Modern international borders
0
50 km
_ Concept Inca Highways
427
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 428
tions were forcibly moved along the road system. To service road traffic, the Inca built a network of roughly 1000–2000 tampus along the roads. The tampus, which tended to be roughly a day’s walk apart, also served as seats of local government. Archaeological studies have found evidence of mining, spinning, pottery manufacture, coca exploitation, and ceremonial activities associated with specific tampu sites. On the eastern Andean slopes of Bolivia, certain fortified sites may have served tampu as well as military functions. Francisco Pizarro’s conquest expedition traveled through Tawantinsuyu along Inca roads, robbed the tampus, and stole the stored food. Ironically, the great road system, which enabled the Inca to control their widespread empire, may also have contributed to their downfall. Figure 9.44 Segment of a 10-mwide Inca road that is still used today by foot travelers.
Communications and their maintenance have been a major concern of Andean peoples from Inca times and probably earlier. In Andean conditions, roads, trails, and bridges were probably as essential to the political cohesion of the Inca state and to the redistribution of goods within it, as they are today to political integration or economic development. —Donald E. Thompson and John V. Murra (1966)
tampu A roadside lodging and stor-
_ _
age place (principally for food, fodder, firewood, and other commodities) along the Inca road system. Tampus were placed roughly one day’s walk apart.
428
wet areas, stone pavements were laid over the road, and drainage canals were built. The Inca are renowned for the several types of bridges they constructed over deep chasms and waterways. Rope suspension bridges, derived from the ancient Andean weaving tradition, greatly impressed the early Spanish chroniclers (Figure 9.45). Some of these bridges, where cables were hung from supporting masonry towers, continued in use into the twentieth century. Other Inca bridges were constructed with superstructures of wood, stone, or even floating reeds. In Inca times, most natives of the Andean highlands could not swim. Therefore, when a bridge could not be built or was being repaired, the Inca generally used ferries of balsa wood or reed. The Inca road system was the lifeline of the Inca state and could not be traveled without an imperial directive. Most often the roads were used to move government messengers, Inca armies, royal litters, or state trade caravans. Sometimes conquered popula-
Chapter 9 South America: The Inca and Their Predecessors
Figure 9.45 A woven Inca bridge over the Apurimac River in Peru, spanning a gap of about 45 m. This engraving was published by E. George Squier, one of the last travelers to see the bridge, in 1877.
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 429
Site
:
Huánuco Pampa A provincial outpost of the Inca
Following his conquest of the Lake Titicaca basin, the Inca Pachakuti began what was to be almost a century of expansion. When possible, Inca ambassadors tried to negotiate a peaceful submission, a strategy that proved successful among smaller polities too weak to defend themselves. Some leaders may have been more willing to join than fight because of the Inca policy that allowed local rulers to retain power as long as they gave their primary allegiance to the Inca. When military action was necessary, the Inca had a large, effective fighting force supplied in part through their road network. Tawantinsuyu, the Inca empire, was divided into four quarters that, according to Spanish chroniclers, were subdivided for administrative purposes into smaller population units. The subdivisions were based on multiples of 10; the waranqa of 1000 taxpayers was the unit most frequently used. The decimal system of administrative divisions was probably coordinated with the quipu, an Andean recording device that uses knotted strings in a positional decimal system (Figure 9.46). As far as we know, neither the Inca nor any of the other prehispanic South American peoples devised a formal system of writing. Quipu, however, are now thought to have encoded more information than just numerical calculations. The patterns of colored strings and knots also may have been able to convey certain narratives. To minimize provincial rebellion, the Inca used a system of colonization called the mitmaq, whereby populations were shifted from one part of the empire to another to break up dissident groups or to improve the exploitation of certain resources. Yet the Inca’s effect on the Andean political landscape was less dramatic than it might seem. A common misconception in popular accounts is that the Inca ruled a mono-
lithic, uniform state. In reality, the empire was an amalgam of many units, which differed in ethnicity, political structure, size, and often language. Inca imperial administration was variable, depending on the economic and political significance of the conquered domain. The Inca tended to make substantial governmental changes only when a region’s pre-Inca organization was fragmentary or when resistance to Inca rule persisted. Spanish chronicles from the sixteenth century emphasize the military and administrative strategies used by the Inca to control their empire. Yet the Inca also established and maintained the authority of the state through a series of civic-ceremonial rituals, as well as through the economic activities on which these ritual events were centered. An archaeological perspective on Inca provincial rule has been illustrated most clearly at the site of Huánuco Pampa (WAHN-u-ko PAHM-pa), a late prehispanic administrative center located in central Peru. In this highland region, the late Craig Morris and Donald Thompson, professor emeritus at the University of Wisconsin, Madison, studied the civic-ceremonial center, as well as the surrounding communities. The Inca rapidly built Huánuco Pampa during the middle of the second half of the fifteenth century A.D. The site was placed on an inhospitable high plain at 3800 m (12,500 ft), an area previously uninhabited and relatively isolated from other settlements. Significantly, Huánuco Pampa lies directly on the main Inca road between Cuzco and Quito. The ruins of the settlement cover roughly 2 sq km (0.8 sq mi). By the time of the Spanish invasion in 1532, Huánuco Pampa included more than 4000 stone buildings (Figure 9.47). Much of this construction is remarkably
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on Pizarro and his relations with the Inca, see the Internet exercises on your online learning center.
Figure 9.46 An official (on the right) uses a quipu to relate accounting information to a noble (on the left).
waranqa A subdivision of the Inca
empire that was used for administrative purposes, consisting of 1000 taxpayers. mitmaq A system of colonization
used by the Inca to minimize provincial rebellion by moving people around to break up dissident groups.
Site Huánuco Pampa
429
_
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Figure 9.47 The plan of Huánuco Pampa; hundreds of warehouses were located on the hillside overlooking the site.
Page 430
N Men’s residences
Cloth manufacture/ women’s residences
Feasting Inca’s compound Public ceremonies
W are
ho
Temporary residences
us
ea
dm
ini
str
ati
on
s
se
ou
eh ar W
Spring
_ _
It seems likely that the city plan of Huánuco Pampa reflected basic Inka organizational principles that permeate several aspects of Inka ideology and practice. The best documented expression of these principles is in the organization of the shrines around Cuzco. . . . There is no written evidence as to the meaning of the complex set of divisions at Huánuco Pampa; however, the similarity to the patterning of the Cuzco shrines and the way they were worshiped seems too close to be coincidental. Similar principles of division and combination pervade other aspects of Inka organization and ideology, such as the calendar. —Craig Morris (1998)
430
0
different from that found in outlying communities. Both the architectural and the ceramic remains at Huánuco Pampa mimic the imperial Inca tradition found at Cuzco, whereas smaller towns and villages retained customary local practices. The Inca stone masonry was finest at the core of the site. An enormous central plaza (550 ⫻ 350 m, or 1800 ⫻ 1100 ft) included a giant stone platform (Figure 9.48) and was flanked by several smaller plazas and large stone compounds (Figure 9.49). The stone-
Chapter 9 South America: The Inca and Their Predecessors
400 m
work in the platform and central compounds was a good imitation of the Inca style, even though it lacked the precise joining of intricately shaped stone blocks seen at Cuzco itself. The extensive storage facilities associated with Huánuco Pampa are indicative of the site’s reliance on its supporting hinterland, as well as the great power of the Inca administration (Figure 9.50). Although Huánuco Pampa’s immediate agricultural hinterland could have supplied small quantities of potatoes, it could not
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
have produced sufficient amounts to feed the site’s permanent residents. Maize found in storage facilities must have been imported from fields at lower elevations. Food supplies and the labor to construct the center must have been drawn from a wide hinterland. Huánuco Pampa was not a manufacturing center and lacked a major market. Much of its populace was an administrative elite, along with their retainers and record keepers. However, in Huánuco Pampa’s central state compounds, Morris and Thompson found unusual concentrations of giant storage vessels associated with the production of maize beer, or chicha, and spindle whorls and other artifacts linked to the manufacture of textiles.
Page 431
Since chicha is not easy to store or to transport, it probably was distributed at Huánuco Pampa during feasts and state rituals. The unusual abundance of large food-preparation vessels and serving platters in the ruins of the central compounds at Huánuco Pampa further suggests the importance of such activities. State-sponsored feasts could have been held in the huge open plaza at the core of the preplanned site, next to the food-preparation and serving areas. Traditionally, the sharing of food and drink was a central feature of prehispanic Andean social relationships. At Huánuco Pampa, customary modes of sharing food and drink probably were used by the Inca state to establish
Figure 9.48 A giant stone platform at Huánuco Pampa.
What role did alcohol consumption play in Inca social organization?
y
chicha A South American beer made
from maize.
Site Huánuco Pampa
_
431
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 432
N
0
5m
Figure 9.49 A set of architectural compounds connected by gateways at Huánuco Pampa. Side View
Overhead View
Solid floor
Exterior surface
Ventilation floor
Figure 9.50 The plan of a two-room rectangular storehouse at Huánuco Pampa.
0
Ventilation floor
432
Duct
Ventilation floor
2m
loyalties and encourage the cooperation of the local population in military, political, and economic activities. The state control and centralization of cloth production also may have played a part. As Murra has noted, textiles had an unusual sociopolitical significance in the prehispanic Andes. Cloth was exchanged at most of life’s major turning points, such as marriage, as well as during major political events, such as conquest and incorporation. The more textiles the Inca state had, the more effectively it could raise and control armies and workers, and the more successfully it could consolidate conquered domains.
_ _
Duct
Solid floor
Chapter 9 South America: The Inca and Their Predecessors
Through the control and manipulation of textiles and administrationsponsored feasting and hospitality, the state was able to cement symbolic ties between the Inca and the conquered people. The extensive roadways, military strength, and innovative bureaucratic technology clearly were critical to the growth and maintenance of the Inca empire. Perhaps critical, too, were the rituals that defined participation in the Inca state as involving more than working state lands, helping build platforms, or fighting in a remote war. Ideological factors also appear to have had a key role in shaping the Inca state.
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
Page 433
Images and Ideas The Organization of State Society Similarities and differences in two prehispanic worlds
T
his chapter’s introduction outlined some of the key similarities and differences between the early civilizations of Mesoamerica and the Andean region of South America. The market, an important institution in prehispanic Mesoamerica, was not nearly as important in the ancient South American world. Whereas the Cortés-led conquerors of the Aztecs were amazed by the size and the splendor of the central Mexican market, the sixteenth-century Spaniards in Peru were impressed not by markets but by the large number of giant storage areas controlled by the Inca state. The Spanish soldier Pedro Cieza de León, writing soon after the fall of the Inca, noted that
The . . . comparative study of the origins of early states has been an important research theme since before the emergence of anthropology as a conscious, distinctive intellectual approach. —Robert Mc. Adams (1966)
in the more than 1200 leagues of coast they ruled they have their representatives and governors, and many lodgings and great storehouses filled with all necessary supplies. This was to provide for their soldiers, for in one of these storehouses there were lances, and in another, darts, and in others, sandals, and in others, the different arms they employed. Likewise, certain buildings were filled with fine clothing, and others, with coarser garments, and others with food and every kind of victuals. When the lord was lodged in his dwellings and his soldiers garrisoned there, nothing from the most important to the most trifling, but could be provided. (1959, pp. 68–69)
The impressive storage facilities, described in the historical accounts, correspond closely to the large storage features associated with such Inca sites as Huánuco Pampa. Researchers conclude, on the basis of written and archaeological evidence, that state-imposed tribute and labor drafts were key to the organization of the Inca, as well as to earlier Andean states. The development of large-scale storage technologies and extensive road systems in Andean South America made possible the later meteoric expansion of the Inca empire, which drew together a remarkable diversity of societies under a single imperial rule. This immense empire, the largest assembled in the prehispanic New World, culminated a long sequence of indigenous strategies through which Andean peoples interconnected the vast environmental and resource mosaic of their world. The naturally fragmented and diverse Andean landscape may have fostered a perpetual search for more productive combinations of goods and resources and for more workable relations with neighboring peoples. In comparison, the largest ancient Mesoamerican political systems were considerably smaller in scale. The cycles of political consolidation and fragmentation ebbed and flowed in both of these global regions. Despite the obvious differences between the ancient urban civilizations that arose in Mesoamerica and South America, the later prehispanic societies that developed in both regions shared a number of important organizational features. We know that even before the beginning of the modern era, the populations of both areas were characterized by differences in wealth and power, which generally became more apparent later in the prehispanic era. These distinctions were manifest in the clothes and ornaments people wore, the food they ate, the variety of goods they had, and the houses in which they lived. For example, in Mesoamerica, Aztec lords, most of whom inherited their social status, dressed in cotton, ate venison, and lived in elaborate residences with retainers and servants. In contrast, the commoners wore clothes made of coarser maguey fibers, ate relatively little animal protein, and lived in small adobe and stone houses. Inca lords ate, dressed, and resided very differently from Andean
< ( : < V : y < p : W
Empires are by nature unwieldy beasts, as difficult to capture descriptively as they are to manage. —Terence N. D’Altroy (2001)
Images and Ideas The Organization of State Society
433
_
pri31057_ch09.pp390-435.qxd
10/1/09
8:17 AM
People sometimes ask me, "Will archaeology survive in the twenty-first century?" If the dramatic discoveries and scientific achievements of the past 50 years are any guide, the answer must be a resounding yes. —Brian Fagan (1998)
Page 434
commoners. Thus in both regions, these peoples, with powerful government institutions and large population centers, also were characterized by a marked degree of social stratification. Nevertheless, we know that personal attributes could be important even in highly stratified societies, such as the Inca, where great warriors or capable administrators could achieve noble status. With the Inca, this social mobility was necessary because their empire expanded so rapidly that there was a nearly constant need to fill administrative posts. Despite social mobility, however, the most powerful ruling positions could in principle be filled only by members of noble lineages. Inca emperors were believed to have ruled by divine right, claiming descent from the sun. Their power was absolute, checked only by the fear of revolt and past custom. Bloodlines were so important in later Inca times that the first wife of the Inca emperor was his full sister. Yet each emperor also had secondary wives, some of whom were the daughters of neighboring rulers. Archaeological research provides the best avenue for understanding how the state and inherited social differences first emerged in human societies. Marked sociopolitical differences were not institutionalized in the earliest human societies, and they seem uncharacteristic of human populations with extensive huntingand-gathering economies. Yet as we have seen in the Andes, inherited inequalities, social stratification, and the state developed long before European contact. Postconquest documentary records of the Spanish can never fully inform us about the ways in which social, political, and economic distinctions first emerged in the Andean world. Even in Mesoamerica, where written records have survived the ravages of time, the beginnings of social and political inequality predate the advent of writing. Many clues may exist at sites such as El Paraíso and Chavín de Huántar. Although they were not the capitals of urban states, these settlements also show indications of emergent inequities in power and wealth. Ultimately, the solutions to these puzzles lie in the archaeological record, yet they will not be achieved quickly or easily. Questions about social inequality and the rise of states, as well as convincing explanations for the different historical pathways that we see for Mesoamerica and South America, are the kind of big, messy research issues that require mountains of carefully collected information, as well as brilliant, yet measured, inferences and ideas. Even then, each question probably has more than one answer or at least complex answers that interdigitate a multiplicity of factors. We can only hope to continue to make the kind of rapid progress in understanding our past that we have seen during the past few decades. Innovative excavations and a renewed dedication to regional settlement pattern surveys in the Andean region have increased our knowledge base. In this chapter, we have seen how scholars are just beginning to unravel the fascinating story of the emergence of hierarchical societies with monumental constructions on the dry Pacific Coast of Peru. These preceramic groups built large mounds and relied heavily on marine resources for their subsistence. Patterns of exchange and communication across diverse environments that later were to become integral to the Inca empire began at that time. Here we have traced the millennial cycles of transition from those early coastal societies through early states (such as the Moche) and larger polities (centered at Tiwanaku and Chan Chan) to the giant empire of the Inca, which broke apart in the face of the Spanish invasion and colonial conquest.
_ _
434
Chapter 9 South America: The Inca and Their Predecessors
pri31057_ch09.pp390-435.qxd
p
10/1/09
8:17 AM
Page 435
DISCUSSION QUESTIONS
w w w. m h h e . c o m / p r i c e i p 6 e
1. How might you explain the monumentality of early sites near the Pacific coast of northwestern South America in the context of apparently limited reliance on domesticated resources?
For more review material and study questions, see the self-quizzes on your online learning center.
2. What role did cloth play in Andean society? 3. In the Andean region, as in many other areas of the world, preindustrial states and empires tended to be relatively short-lived. Can you explain their fragility? 4. What social and economic factors integrated the Aztec and Inca empires? Compare and contrast.
z
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e.
Bruhns, K. O. 1994. Ancient South America. Cambridge: Cambridge University Press. A synthesis of archaeological research on ancient South America. Burger, R. L. 1992. Chavín and the origins of Andean civilization. London: Thames & Hudson. An illustrated discussion of Chavín de Huántar and earlier Andean sites. Burger, R. L., and L. C. Salazar. 2004. Machu Picchu: Unveiling the mystery of the Incas. New Haven, CT: Yale University Press. A well-illustrated companion to a recent museum exhibition; includes current interpretations of this incredible site. Isbell, W. H., and H. Silverman, eds. 2002. Andean archaeology I and II. New York: Kluwer Academic/Plenum.
A two-volume set of scholarly essays that focuses on new interpretations based on recent archaeological and ethnohistorical fieldwork in the central Andes. Kolata, A. L. 1993. The Tiwanaku: Portrait of an Andean civilization. Cambridge: Blackwell. A synthetic overview of Tiwanaku and its predecessors. Moseley, M. E. 2001. The Incas and their ancestors: The archaeology of Peru, rev. ed. London: Thames & Hudson. A readable overview of Andean prehistory. Stanish, C. 2003. Ancient Tiwanaku: The evolution of complex society in southern Peru and northern Bolivia. Berkeley: University of California Press. A recent scholarly overview of this key Andean region.
_ Suggested Readings
435
pri31057_ch10.pp436-505.qxd
10/1/09
C
8:21 AM
H
A
Page 436
10 P
T
E
R
T
E
N
Pharaoh
Priests and nobility
Scribes
Artisans
Soldiers
Merchants
Peasants and slaves
_ _
Figure 10.1
Schematic representation of the hierarchical tiers of socioeconomic groups in the ancient Egyptian world.
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 437
< States and Empires( in Asia and Africa : Introduction < V : I y < p : W Asia and Africa after the Transition to Agriculture The rise of early states and urban centers
The study of man and civilization is not only a matter of scientific interest, but at once passes into the practical business of life. We have in it the means of understanding our own lives and our place in the world, vaguely and imperfectly it is true, but at any rate more clearly than any former generation. The knowledge of man’s course of life, from the remote past to the present, will not only help us to forecast the future, but may guide us in our duty of leaving the world better than we found it. (Tylor, 1960, p. 275)
n many ways, E. B. Tylor’s words are as meaningful today as they were when first published in 1871. With more than a century of archaeology behind us, we can describe the trajectories of long-term change for a number of world regions, as well as compare and contrast the sequences from those areas. Although a thorough understanding of these differences and similarities remains ahead of us, today we know much more and therefore can ask research questions that are more specific than was possible in Tylor’s day. The schematic picture of different segments of ancient Egyptian society (Figure 10.1) illustrates the nature of state societies past and present. As shown in this graphic image, such hierarchical social formations are composed of different societal segments and socioeconomic strata, each associated with different professions and economic roles. The specific people depicted provide clues to some of the broad societal realms—economic, military, and ideological—that form the bases of power for the leaders and governing authorities at the top of the social pyramid. Of course, the precise mix of (and interconnections between) these foundations or underpinnings of power historically varies both in time and across space, as does the manner in which power and decision making were wielded by those at the summit of the sociopolitical hierarchy. As in other preindustrial states, this picture of ancient Egypt depicts the rulers at the apex of the earthly universe, closer to the heavens and supernatural forces, from whence they could shape important aspects of belief systems and ideology. The historical factors and underlying causes that underpinned the emergence of hierarchical social formations also varied greatly. Nevertheless, in many regions of the world, including Mesoamerica (Chapter 8) and Andean South America (Chapter 9), long-term shifts did result in the rise of polities and institutions that were hierarchically organized. Such institutions emerged in conjunction with
437
CHAPTER OUTLINE
Introduction Asia and Africa after the Transition to Agriculture 437 SITE Eridu
441
CONCEPT Temples SITE Uruk
445
446
CONCEPT Early Writing Systems
451
SITE Harappa and
Mohenjo-daro 454
CONCEPT Economic Specialization 460 SITE Hierakonpolis
462
CONCEPT The Cemetery at Hierakonpolis 467
SITE Giza and Dynastic Egypt CONCEPT Pyramids SITE An-yang
468
473
475
CONCEPT The Roots of Chinese
Cuisine
480
SITE Xianyang SITE Angkor
482
488
SITE Jenné-jeno
492
SITE Great Zimbabwe
496
Images and Ideas Theories of State Development 501
_
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
w w w. m h h e . c o m / p r i c e i p 6 e
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
Power relations are demonstrably the most complicated and most important aspect of the governance of human societies. It is particularly important to know how power shifts occur and under what conditions various power distributions constitute stable and unstable configurations. . . . [There is] great utility in the study of change and [to] be able to explain how certain forms of governance can be associated with particular histories of environmental stability and instability. This gives equal value to all social formations, inasmuch as we can learn important lessons from them all. —Carole L. Crumley (1995)
Mesopotamia The flat plain between
_ _
the Tigris and Euphrates rivers in southern Iraq where the world’s first civilization developed.
438
Page 438
larger, denser populations, where consensual methods of decision making no longer work effectively. At the same time, hierarchical institutions, characterized by differential power and authority, tend to become self-serving as well as systemserving. They work to perpetuate and enhance the extant structure (and its inequalities) so that those who are already advantaged remain the privileged class. Increasing inequality in wealth and power is a hallmark of so-called civilized societies. Such stratification also may be related to the problem of succession that characterizes all hierarchical institutions. By specially marking and treating the offspring of the privileged from birth, society can expose them to the knowledge, personal connections, and traditions necessary for assuming power. At the same time, such marking can limit the number of rightful heirs, thereby minimizing (if not eliminating) the disruptive consequences of succession disputes. Of course, wealth differentials and stratification are associated with the unequal control of power, land, and resources. As we see in this chapter, the rise of hierarchical polities is generally linked with economic transitions in exchange and production. Particularly in Eurasia and Africa, where beasts of burden and wheeled vehicles could distribute goods relatively inexpensively, we see the development of increasingly large-scale and specialized craft industries. Although such industries frequently were controlled by independent entrepreneurs, over time they offer central political institutions greater opportunities for limiting access and concentrating wealth. As societies increase in size and organizational complexity, the mechanisms of exchange also shift from face-to-face contacts to tribute and marketing. Although these general trends are evident for the regions discussed in this chapter (Figure 10.2), the specific sequences of transition and the rates of change were not uniform. Likewise, the specific organization of different states varies markedly. Some were focused on the exclusive power of individual rulers or families; others had more dispersed power-sharing arrangements. The world’s first states evolved in Mesopotamia, where the temple institution became a key focus. (For information on the earlier antecedents of topics and sites highlighted in this chapter, see discussions in Chapter 6.) We begin our discussion with Eridu, the site with perhaps the earliest Mesopotamian temple, established by the end of the sixth millennium B.C. The Mesopotamian temple functioned differently than the Mesoamerican temple. The former had a key economic role, not seen in the Mesoamerican institution, in that it received goods through tribute and then redistributed a portion during feasts and other activities. We next examine the early Mesopotamian urban center of Uruk, situated amid a network of ancient canals not far from the Euphrates River. Scholars agree that large-scale canal irrigation was a key feature of early Mesopotamian civilization, and the remnants of the ancient waterways are still evident on the desiccated landscape. However, the specific causal relationship between water management and state development in Southwest Asia remains a matter for debate. Some argue that the allocation of water and the maintenance of the canals necessitated some kind of central authority; others suggest that the monumental water-control systems were built only after the rise of powerful states. Whichever scenario is eventually supported, later empires of Southwest Asia capitalized on the great grain-producing potential of large-scale canal irrigation systems. Slightly later, during the third millennium B.C., major centers arose along the Indus River and its tributaries in what are today Pakistan and India. Here we focus on the best-known sites of the Indus civilization: Harappa and Mohenjodaro. Although these South Asian sites have not yielded rich tombs like those in Mesopotamia, they are known for their highly developed craft industries. Indus centers generally were not as large as those in Mesopotamia, but they appear to have been more systematically planned, with centralized drainage networks for individual houses.
Chapter 10 States and Empires in Asia and Africa
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 439
To a degree, the rise of civilization in Egypt was inspired by political and economic ties with Southwest Asia. Yet the course of development was different. Large-scale political centralization was much more in evidence in Egypt than in Mesopotamia. Power was more focused on the principal ruler (viewed as a divine king) in Egypt. The monumental pyramid tombs of Egypt represent an expression of wealth and grandeur unsurpassed in ancient Southwest Asia. For Egypt, we begin with the center of Hierakonpolis, a major settlement along the Nile whose occupation largely predates the unification of northern and southern Egypt. We next discuss the later funerary complex at Giza, located close to the ancient capital at Memphis. Both were important places in Dynastic Egypt. At Giza, several centuries after unification, powerful pharaohs constructed some of the world’s largest pyramids. By early in the second millennium B.C., states had developed indigenously in North China. An-yang, the last capital of the Shang dynasty, was excavated first in the late 1920s and remains one of the best-known early Chinese cities. Although early Chinese centers were somewhat less urban than those in early Mesopotamia, they became some of the largest in the world by A.D. 1. During the last centuries B.C., in the reign of the Qin dynasty, a magnificent tomb was built near the capital of Xianyang. The mausoleum included a huge army of life-size terracotta (brown-orange earthenware) foot soldiers and cavalrymen and illustrates the extreme stratification in ancient China. After the rise of early Asian civilizations, waves of communication and exchange by land and sea between the Indus region, China, and Southeast Asia helped fuel waves of political expansion and fragmentation in Thailand, Cambodia, Vietnam, Laos, and neighboring lands. The Angkor state that arose in Cambodia just after A.D. 800 was one of the largest and most centralized of these Southeast Asian polities, and the Khmer rulers who reigned there endured for more than 600 years. Although they borrowed from South Asian belief systems, the Khmer’s distinctive form of kingship was markedly different from that of the much earlier Indus structure of rule, which did not emphasize specific individuals. Rather, one can draw parallels between the Khmer kings and the noble lords of Classic period Maya society (who were entirely separate historically and half a world away). As with the Classic Maya, Khmer kingship emphasized their divine monarchy and fostered the concentration of wealth and luxury in their hands. Finally, we discuss sub-Saharan state development in western Africa at the tell site of Jenné-jeno and in southern Africa at Great Zimbabwe. The early urban center of Jenné-jeno was an important node in long-distance trade between the Sahara and gold-producing areas to the south during the first millennium A.D. Although not the earliest state in southern Africa, Great Zimbabwe is a spectacular site with large stone structures erected early in the second millennium A.D. In the chapter’s summary, “Images and Ideas,” we return to some of the key models and ideas that have been advanced to account for early state development. In a world that often seems to be dominated by powerful presidents, despots, kings, legislatures, bureaucracies, and tax collectors, it is important to understand how such hierarchical systems arose in the first place and why individual households lost key elements of their autonomy. To answer such questions about the beginnings of inequality and the rise of the first hierarchical decision-making institutions, we must turn to archaeology. In most parts of the world, the processes that culminated in unequal access to power and wealth began centuries before the advent of written records. As a consequence, written histories alone cannot answer these evolutionary questions. Archaeology must continue to focus on the dramatic changes that occurred between the advent of agriculture and the rise of urbanism in many areas of the world. Unraveling the processes and causal connections surrounding those episodes may help us explain the course of recent human history.
Interpreting the relationship between people and their environment for developing future scenarios of people on Earth requires, as a first step, an understanding of human history beyond current accounts, which are mostly informed by the political ideologies of the “modern” nation-state. —Fekri Hassan (2007)
terracotta A hard, brown-orange earthenware clay of fine quality, often used for architectural decorations, figurines, and so on.
_ Introduction Asia and Africa After the Transition to Agriculture
439
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 440
ATLANTIC
An-yang
OCEAN Uruk Giza
Xianyang Eridu
Harappa Mohenjo-daro
Heirakonpolis
PACIFIC OCEAN Angkor
Jenné-jeno
Equator INDIAN OCEAN Great Zimbabwe
• Sites emphasized in this chapter
Eridu Uruk Hierakonpolis Giza Harappa and Mohenjo-daro An-yang
site at its largest extent
Xianyang Angkor Jenné-jeno Great Zimbabwe
episode of occupation
6000
5000
4000
3000
2000 Year
_ _
Figure 10.2 Location of and timeline for sites in Asia and Africa.
440
Chapter 10 States and Empires in Asia and Africa
1000
B.C./A.D.
1000
Present
pri31057_ch10.pp436-505.qxd
10/15/09
11:07 AM
Page 441
Site
:
Eridu An early ceremonial center in Mesopotamia
TARU
S MOUNTAINS
TURKEY
Caspian Sea
Tell Hamoukar Tepe Gawra
Tell Brak Nineveh
SYRIA Z
ASSYRIA
A
G
RO
e rat ph Eu
Tell as-Sawwan
IRAN
S M
.
Baghdad Ti gr is R.
NT AI NS
ANCIENT COUNTRY
OU
R
s
For nearly 10,000 years, the people of Southwest Asia have lived in farming communities. Their settlements, consisting of small, closely packed mud or clay rectangular structures, are ideally suited to the climate and the available resources. But mud-brick dwellings deteriorate rapidly. They must be rebuilt every 50–75 years, after a few generations of use. Over millennia, mud-brick abodes have been rebuilt repeatedly on top of earlier structures. Gradually, large mounds of accumulated mud and clay, known as tells, form across an entire community. Thousands of tells rise above the landscape of Southwest Asia, some as high as 50 m (165 ft) above the surrounding terrain. These mounds contain debris from thousands of years of habitation. Ancient irrigation canals also cover the landscape of Southwest Asia. For the past 7000 years, irrigation has been practiced in this part of the world, and remnants of canals remain a visible reminder of the region’s rich archaeological past. The soils of the alluvial plain are deposited by the annual floods of the Tigris and Euphrates rivers. The rivers also provide the water that makes irrigation (and hence agriculture) possible in this region, where rainfall is inadequate for farming. Today, Mesopotamian farmers using irrigation can cultivate a variety of crops, including wheat, barley, dates, lentils, olives, oranges, and onions. Along the swamps formed by the rivers, several kinds of usable plants, such as flax for textiles and rushes for basketry, can be collected. Although scrub forests along the rivers do not support much game, fish are abundant in the rivers. However, the riverine lowlands of the Mesopotamian alluvium do not contain much in the way of raw materials.
MODERN COUNTRY Archaeological sites Modern cities Modern international borders
SUM ER Uruk
0
100 km
Ur Eridu
Basra
IRAQ
SAUDI ARABIA
No early agricultural villages, such as those established in the coastal Levant and in the foothills of the Taurus and Zagros Mountains (see Chapter 6), have been recorded on the flat plain between the Tigris and Euphrates rivers in southern Iraq, the area known as Mesopotamia. The alluvial floodplains and deltas here are hot and dry, composed primarily of inhospitable sand, swamp, and dry mudflats. Yet it was in Mesopotamia that the world’s first civilization developed (Figure 10.3). By 6000 B.C., the first scattered farming villages were settled on the northern fringe of the Tigris and Euphrates floodplain, an area where seasonal rainfall
KUWAIT
Persian Gulf
Figure 10.3 Mesopotamia, or the land between two rivers (the Tigris and the Euphrates), and sites mentioned in the text.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on medicine in ancient Mesopotamia, see the Internet exercises on your online learning center.
_ Site Eridu
441
pri31057_ch10.pp436-505.qxd
10/15/09
Figure 10.4 The remains of the large stepped temple mound at Eridu.
monochrome One color; describing
_ _
pottery decorated with only one color that contrasts with the underlying color of the paste of the vessel.
442
11:05 AM
Page 442
can sustain agriculture. These communities were generally composed of several houses, entered through the roof, containing two or three rooms each. Ovens and chimneys were common features. By 5500 B.C., a new style of painted pottery, called Halafian, had spread throughout a wide area of northern Mesopotamia, replacing the monochrome wares that had been made previously. The remarkable similarity of Halafian pottery over a wide area suggests that small villages were linked into far-reaching networks. Shortly after the appearance of Halafian ceramics in the north, the focus of Mesopotamian settlement shifted somewhat to the southern alluvial plain of the Tigris and Euphrates rivers, an area known as Sumer. No sedentary villages before the sixth millennium B.C. have yet been recorded in this area. That is not surprising, because the area lacks plants and animals that later were domesticated, and even usable stone. In most years, the region receives insufficient precipitation for dry farming. Yet with an economy based on fish, irrigation agriculture, and domestic cattle, Sumer became the demographic and political core of
Chapter 10 States and Empires in Asia and Africa
Southwest Asia for most of the next 4000 years. The ’Ubaid (oo-BAID) period (c. 5300–4100 B.C.), which begins this sequence, was marked by an increasing reliance on canal irrigation and the establishment of the temple (see “Temples,” p. 445). The first villages in southern Mesopotamia were small, mostly 1–2 ha (2.5– 5 acres). Yet, relatively rapidly, a few of the settlements, such as Eridu (AIRih-dew), grew in size and importance. Eridu, the earliest known settlement on the southern Mesopotamian alluvium, was established before the end of the sixth millennium B.C. Ancient Sumerian accounts of creation (written in 3000 B.C.) name Eridu as one of the first communities to emerge from the primeval sea. Although most of the structures at Eridu were houses, the initial occupational levels feature a significant nonresidential structure. The architectural plan of this public building bears sufficient similarity to later Sumerian temples (depicted in written texts) to suggest that it may have served as an early temple. The building, constructed of mud bricks, measured 3.5 ⫻ 4.5 m (11.5 ⫻ 15 ft), with a possible altar facing the
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
entrance and a pedestal in the center of the room. Signs of burning on the pedestal indicate that it may have been used for offerings. This possible temple suggests that the organizational capacity for the construction of minor public architecture was present; however, its small size reflects a much lower level of social complexity than in later times. Yet the temple institution may have had antecedents that preceded the movement of populations into Mesopotamia’s southern alluvium. An earlier T-shaped structure at Tell asSawwan, on the border between the northern plain and the southern alluvium, may have been a public building with functions similar to those of the later temple. In addition, a circular domed structure at Tepe Gawra, on the fringes of the northern plain, was found below a sequence of superimposed temples. At other northern Mesopotamian sites, similar circular structures, often joined to a rectangular entryway and thereby making a keyhole shape, have been found. These structures, or tholoi, were distinct from the traditional rectangular dwellings and seem to have had a nonresidential purpose, perhaps related to ritual and storage. Because the temple institution was a focal point of early civilizations in Southwest Asia, questions about its origin and antecedents are important for Southwest Asian prehistory. Yet there also are more general implications. If the temple emerged only in Sumer, its development may be linked to the increasing and necessary reliance on canal irrigation. Social theorists have long argued that the management of irrigation systems requires cooperation among farming populations to allocate water and maintain canals—a particular problem in Sumer, where the rivers carry and deposit great quantities of silt. Alternatively, if antecedents of the temple were established before the occupation of the southern Mesopotamian alluvium, perhaps the movement into Sumer became possible only once a central redistributive institution,
Page 443
such as the temple, was in place. Without that integrative institution, the agricultural hazards of flooding, drought, and dust storms could not have been overcome. In this scenario, irrigation management may have been responsible for the expansion and elaboration of the temple institution but not for its initiation. Likewise, another consequence of irrigation is that it tends to enhance disparities in agricultural productivity and hence land value. In Sumer, emerging inequalities in agrarian production may have fostered increasing economic stratification. Much more archaeological research is needed to address these issues adequately. By 4500 B.C., the southern Mesopotamian alluvium was dotted with full-fledged towns and public buildings (Figure 10.4). Based on irrigation farming, the economy produced enough food to support a growing population, yielding a surplus that supported craft producers and decision makers. Eridu may have covered 10 ha (25 acres) by that time and had a population as large as several thousand people. The temple at Eridu was rebuilt numerous times and expanded so that it contained multiple altars and offering places (see Figure 10.7). The most elaborate residential dwellings were situated immediately around the temple at the center of the community; craftspeople and peasants lived at ever-increasing distances from the core of the settlement. The development of the temple institution and the spread of canal irrigation were key features of the ’Ubaid period in Mesopotamia. This period was identified by a widespread monochrome pottery decorated with geometric designs (Figure 10.5). ’Ubaid times also were characterized by population growth and increases in craftwork. ’Ubaid pottery was made in a wider range of forms than the earlier Halafian ware. Yet it tended to be somewhat less decorated, and the different ceramic varieties were generally more uniformly distributed over space. Most of the ’Ubaid ceramics appear to have been made on a slow-turning potter’s wheel, in use for the first time.
0
10 cm
Figure 10.5 Two examples of ’Ubaid pottery decorated with small rectilinear patterns. Other common designs include triangles, grids, and zigzag lines.
tholoi Ancient Mesopotamian round
structures that often were attached to a rectangular antechamber or annex, resulting in a keyhole shape.
_ Site Eridu
443
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 444
Figure 10.6 An artist’s reconstruction of the Uruk period temple at Eridu.
Perhaps the most significant theoretical trend to have emerged in the last 15 years is the fact that anthropological archaeologists have shifted their focus away from attempts to generate and test explanatory models for the origins of primary or pristine states in the Old World. Instead, we now see an emphasis on “state dynamics”—how these societies actually functioned. With only a few exceptions, the current studies of state dynamics emphasize the specific factors that account for variation in the ways these polities were organized. —Gil Stein (2001)
_ _
seal stamp A piece of inscribed stone used by administrators to impress a symbol on wet pieces of clay or bitumen in order to keep track of goods.
444
Because of the absence of suitable sources of stone, hard-fired clay was used to make sickles, mullers (implements used to grind paints, powders, etc.), hammers, and axes. The complex processes that led to the growth of later civilization in Mesopotamia clearly had begun during the ’Ubaid period. Yet indicators of pronounced social differentiation appeared only at the very end of the period. There are few exotic luxury items at ’Ubaid sites. And despite many excavated ’Ubaid burials, no highly elaborate funerary contexts have been unearthed. Not until the subsequent Uruk period (4100–3100 B.C.) did monumental urban centers arise in accordance with clear indicators of social stratification. Eridu remained an important place for more than 1000 years following the end of the ’Ubaid period (Figure 10.6). Yet early in the fourth millennium B.C., other centers, such as Uruk (also known as Warka), rapidly surpassed it in size, monumentality, and political significance. Although for much of its early history the centers and polities of Mesopotamia shared a common cultural tradition, rarely was this region dominated politically by a single ruler or core state. Recently, important findings have been made at the late fifth and early fourth millennium occupations of Tell Hamoukar and Tell Brak in northern Mesopotamia. At these sites, deeply
Chapter 10 States and Empires in Asia and Africa
buried under later deposits, indicators of early Mesopotamian urban life, including seal stamps used for impressing symbols on wet pieces of clay, were recovered. Seal stamps have been commonly found at Uruk and other sites in southern Mesopotamia. Their presence, along with evidence of early monumental constructions and fortification walls, indicates that early Mesopotamian urbanism was taking root in both the northern and southern parts of this region. It also is important to note that archaeologists are not merely concerned with the rise of states. In recent decades, they have expanded their examination into the long-term dynamics of polities in a given region over time. In studying these dynamics, it also has proven important to investigate how the different components of archaic states worked or were interconnected. Such research is an important basis from which to explore transitions and/or episodes of collapse.
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 445
Concept
<
Temples The role of the temple institution in Mesopotamia
Temples were established in southern Mesopotamia no later than the ’Ubaid period (c. 5300–4100 B.C.), and by the end of the ’Ubaid, the Sumerian pattern of towns with temples was well entrenched. In later Mesopotamian cities, the temple was always the largest and most impressive building, and it had both economic and religious functions. The temple institution became the focal point for the powerful religion and statecraft that integrated and maintained the complex polities that arose in ancient Mesopotamia. At Eridu, a series of about 20 superimposed temples, spanning more than 3000 years, was excavated (Figure 10.7). Built atop earlier ruins, the later ones were raised above the land. The remains of earlier shrines are preserved within the foundations of later buildings. Each structure was separated by deposits of debris, including small animal and fish bones, that may have been offerings. Although not all temples are associated with great stepped towers known as ziggurats, many are. How significant the ziggurats were is evident in the names by which they are known—House of the Mountain, for example. In Mesopotamia, the word mountain has great religious significance, as the place where the power of all natural life is concentrated. In a terracotta relief found in a temple from the second millennium B.C., the body of the deity issues from a mountain. For Mesopotamia, third-millennium texts tell us that the laws of the gods were unchanging and people were governed by the gods’ decisions. But the rapid rise in the importance of the temples cannot be explained fully by religion. Their growth, and the cities that developed around them, was closely related to the economy. Generally associated with storage facilities, the temple became a redistributive center for agri-
Third-dynasty ziggurat
I
V VI II III IV VII
Present ground level
cultural produce and craft goods. It also provided help after floods, dust storms, and scorching winds when crops failed. Temple administrators organized the cooperative projects necessary to construct and maintain irrigation channels and probably decided who received how much water. Canal construction created unequal land values and inequities in land holdings. The temple elite owned land, employed people directly, and were involved in farming and manufacturing activities. Because they controlled large quantities of food, they could support full-time craft specialists, including stonemasons, weavers, and copperworkers. The temple also directed the long-distance trade of raw materials for the craft specialists. Status goods were received in return, reinforcing the position of the temple elite. The temple became the central economic force. Building ziggurats required huge amounts of labor and materials. That so much energy was diverted to nonutilitarian tasks attests to the power of the temple elite. Monumental architecture served to reinforce social and political hierarchies. Temples may have been objects of civic pride, while also validating the power and authority of the ruling elite through rituals and religious sanctions.
VIII IX XI
XV
XVI
Figure 10.7 The long sequence of temples excavated at Eridu, with the sequence of construction indicated by Roman numerals. The earliest temple (XVI) was very small in comparison with the increasing size and elaboration of later temples.
ziggurat A large pyramid in
Mesopotamia consisting of many stepped levels.
Concept Temples
_
445
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 446
Site
:
Uruk The world’s first monumental urban center
[T]he primacy of southern Mesopotamia was in part due to the fact that southern societies had several important material advantages over polities in neighboring areas. These included (1) a denser and more varied concentration of exploitable subsistence resources, (2) higher and more reliable agricultural yields, and (3) an exponentially more efficient distribution system based on water transport. These advantages promoted the creation of inherently asymmetrical exchange patterns among independent polities in the Mesopotamian alluvium and between those polities and societies in neighboring regions which, over time, produced important organizational asymmetries between southern societies and contemporary polities. —Guillermo Algaze (2001)
Many of the processes underlying the growth of urban civilization in Sumer began during the ’Ubaid period, but it was not until the succeeding Uruk (4100–3100 B.C.) and Early Dynastic (3100–2370 B.C.) periods that urban settlements and the earliest states were first established. Based on the great surplus potential of intensive irriga-
tion agriculture, the power of southern Mesopotamian cities and their rulers eventually came to exceed that of those in the north. In Mesopotamia, the site of Uruk (oo-ROOK), with its giant stepped pyramid, the Anu Ziggurat, was the largest and most impressive. During the fourth millennium B.C., Mesopotamia also entered the era of
North gate
City wall
Anu Ziggurat
City wall
N
_ _
Figure 10.8 The plan of Uruk, showing the principal platforms and ziggurats and the 5000-year-old wall that encircled the site.
446
0
500 m
Chapter 10 States and Empires in Asia and Africa
Ur gate
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
history, with the use of a system of writing on clay tablets. Located on the banks of the Euphrates River less than 160 km (100 mi) north of Eridu, Uruk was settled before 4000 B.C. during the ’Ubaid period and became a major city of more than 10,000 people, covering 100 ha (250 acres) by 3100 B.C. (Figure 10.8). Residential structures at Uruk were made of whitewashed mud bricks. Houses were rectangular and were built along narrow, winding streets. Most outlying sites were small, only 1–3 ha (2.5–7.5 acres) in size (smaller than a city block), although several settlements of intermediate size were established. Rapid population growth during the Uruk period was probably due to in-migration and perhaps the settling down of nomadic peoples, in addition to natural increases resulting from changes in fertility and mortality rates. The earliest monumental architecture at Uruk, the Anu Ziggurat, is composed of a series of building levels, the earliest going back to ’Ubaid times. This stepped pyramid, named for the primary god (Anu) in the Sumerian pantheon, attained its maximum size at the end of the Uruk period. At that time, the White Temple was built on top of the Anu Ziggurat, 12 m (40 ft) above the ground. The temple, which measured 17.5 ⫻ 22.3 m (57.5 ⫻ 73 ft), was made of whitewashed mud bricks and decorated with elaborate recesses, columns, and buttresses. Inside, the pattern followed the tripartite temple plan described for Eridu centuries earlier, which consisted of a long central room with a row of smaller rooms on each side. The Anu Ziggurat and the White Temple mark a transition that was occurring in Mesopotamian society. The earlier Eridu temples were small compared with Uruk temples and probably were administered by civic-ceremonial functionaries who had only limited influence over the populace. However, the economic and political power of the temple expanded as the size of its buildings increased. The Anu complex, estimated to have taken 7500
Page 447
man-years to build, represents the control of a large, organized labor force. As has been argued for medieval Europe, “What constituted the real basis of wealth . . . was not ownership of land but power over men, however wretched their condition” (Duby, 1974, p. 13). As temples became more elaborate, the structures, as well as the individuals associated with them, were separated from the general populace. Platforms and ziggurats elevated the temples above the rest of the community. In the Early Dynastic period, temples and the associated priestly residences were often enclosed by high walls that further divided and protected this increasingly wealthy precinct from the lower socioeconomic strata below. By the middle of the fourth millennium B.C., economic specialization in a wide variety of arts and crafts was evident, including stonework (Figure 10.9), metallurgy, and pottery production. Earlier Halafian and ’Ubaid potters had crafted beautifully painted vessels, but during Uruk times, these wares were largely replaced by unpainted pottery (a trend that may have begun earlier in the ’Ubaid period). An important Uruk development was the widespread use of the potter’s wheel, which allowed for the more rapid production of fired clay containers. Using the wheel and molds, potters began to produce mostly undecorated utilitarian vessels in great volume, including crudely made beveled-rim bowls that may have been used for rations (Figure 10.10). In earlier Halafian and ’Ubaid times, highly decorated pottery vessels may have served as wealth or status items that also marked or defined certain social, territorial, or kin groups. By the Uruk period, ceramic items appear not to have played those roles, since they were even plainer and more uniformly distributed over space. Perhaps textiles and metal items largely replaced pottery as wealth items and markers of social identity. As craft specialization and the demand for raw materials increased,
pantheon The officially recognized gods of a people.
_ Site Uruk
447
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 448
Figure 10.9 The Warka vase, an elaborately carved stone vessel on a pedestal, standing about 1 m (3.3 ft) high. The three registers record scenes of domestic and religious life: The lowest register depicts plants and animals; the middle register is a procession of males carrying offerings of food and wine; and the top register portrays a ritual in which a goddess is presented with offerings of food.
The Mesopotamian use of bread in brewing has led to much debate among archaeologists, some of whom have suggested that bread must therefore be an offshoot of beer making, while others have argued that bread came first and was subsequently used as in ingredient in beer. It seems most likely, however, that both bread and beer were derived from gruel. A thick gruel could be baked in the sun or on a hot stone to make flatbread; a thin gruel could be left to ferment into beer. The two were different sides of the same coin: Bread was solid beer, and beer was liquid bread. —Tom Standage (2005)
_ _
carnelian A red or reddish variety of chalcedony (a translucent variety of quartz) used in jewelry.
448
trade flourished along major waterways, especially the slow-moving Euphrates. Ships sailed up the rivers from the Persian Gulf, carrying food and raw materials, including shell, carnelian, silver, gold, lapis lazuli, onyx, alabaster, ivory, textiles, timber, and skins. Copper, which first appeared on the plateau to the north as early as the fifth or sixth millennium B.C., was imported into lower Mesopotamia around 3500 B.C. Metal implements soon took on an important role
Chapter 10 States and Empires in Asia and Africa
in agriculture and warfare. Coppersmiths were present in most Mesopotamian cities by 3000 B.C. The wheel was introduced during the fourth millennium B.C., and wheeled vehicles, drawn by horses, asses, and oxen, became vital in trade and warfare. Another important development during the fourth millennium B.C. was the invention of the plow, which resulted in increased agricultural yield. In large part, Uruk’s power depended on the city’s ability to extract agricultural surplus from the hinterlands. In the southern Mesopotamian alluvium, agricultural surplus was vital, because the region lacked mineral and stone resources. The major crops were wheat, barley, flax, and dates; cattle raising and fishing also were important. The world’s earliest known written documents, clay tablets dating to 3400 B.C., come from Uruk (Figure 10.11). The principal function of this earliest Mesopotamian writing appears to have been economic; the clay tablets record lists of commodities and business transactions. Over 1500 symbols have been identified in these early texts. Signs for carpenter, smith, chariot, copper (in the form of an ingot), plow, and harp have been recognized. Lists of commodities include dairy products, cattle, wheat, barley, bread, beer, clothing, and flocks of sheep. The elements of the early Uruk writing system have been shown to be the forerunners of later Sumerian cuneiform (see “Early Writing Systems,” p. 451). By the beginning of the Early Dynastic period (about 3000 B.C.), Southwest Asian civilization was well established. Written records enable us to trace the dynastic succession of specific kings. Metal tools, which were far more efficient than earlier tools, became much more common. Smiths began to alloy copper with tin to produce bronze, which is much harder than copper. The development of bronze weapons was directly linked to the increasing role of warfare as a means of attaining political ends. Armies were equipped with wheeled chariots and wagons. Rulers became more despotic, concentrating
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
wealth and controlling subjects by military strength, religious sanction, and taxation. During the Early Dynastic, Sumer was divided into 10–15 contemporaneous city-states, which were largely politically autonomous. Uruk grew to 400 ha (1000 acres) and may have contained as many as 50,000 people. This demographic growth was indicative of the great nucleation (in cities) that characterized Early Dynastic Mesopotamia. As the wealth and power of these closely packed cities increased, so did the competition between them. Great defensive walls were constructed around the major urban centers, including Uruk. This period also was characterized by dynastic rule, as individual monarchs and their courts gained increasing independence from long-powerful temple institutions. After its peak, around 2700 B.C., Uruk’s supremacy was challenged by other early cities, and its political importance eventually declined. Yet no single city-state dominated the Mesopotamian landscape for long during the Early Dynastic. Ur, a smaller center just 120 km (75 mi) away, became Uruk’s economic and military rival. Ur was inhabited as early as the Uruk period, but the settlement rose to great prominence only during the subsequent millennium. Early Dynastic Ur is renowned for the Royal Cemetery, which was excavated in the late 1920s by British archaeologist Sir Leonard Woolley. The scale of these excavations is one that we would not undertake today (Figure 10.12). More than 2500 burials were unearthed, providing graphic evidence for superb craftsmanship, opulent wealth, and developed social stratification. Fewer than 20 graves actually contained royal individuals, who were placed in private chambers made of stone blocks and mud bricks. The contents of these graves indicated great concentrations of wealth and the trappings of earthly power. Perhaps the best-known vault is thought to contain the body of Queen Shub-ad. She was lying on her back on
Page 449
a bed, accompanied by female attendants. Two wagons drawn by oxen and attended by male servants had been backed down the entry ramp, where 59 bodies, mostly female, were on the ground near the tomb chambers. All retainers were lavishly bedecked with
Figure 10.10 Examples of undecorated, crudely made beveled-rim bowls. These vessels were made quickly, using molds, and likely were employed for rations, an interpretation supported by Late Uruk texts in which the sign for “to eat” includes a vessel resembling a beveled-rim bowl.
Figure 10.11 A clay tablet from Uruk, approximately 6 cm (2.4 in) wide, records a list of commodities and economic transactions.
_ Site Uruk
449
pri31057_ch10.pp436-505.qxd
10/1/09
Figure 10.12 Sequence of photographs showing the progress of Woolley’s excavations at Ur in 1933–1934.
8:21 AM
Page 450
crafted ornaments made of gold, silver, carnelian, lapis lazuli, and turquoise. Woolley believed that all the people and animals buried with the queen entered the vault alive. After the queen and her possessions were placed in the pit, the animals were dispatched by their keepers, who then consumed poison that was ready for them in the shafts. No violence or confrontation is evident in the arrangement of the corpses. Although royal interments are few at Ur, numerous graves contain modest quantities of goods, and an even larger number include little or no material wealth. Great disparities in individuals’ treatment at death are evident at Ur, suggesting that social stratification was marked. By the end of the Early Dynastic period, bureaucratic organization, so-
_ _
450
Chapter 10 States and Empires in Asia and Africa
cial stratification, trade, crafts, and writing were all highly developed. Yet, if anything, the pace of military conflict and political upheaval was intensified. For Sumer, the history of the third and second millennia B.C. (and beyond) is extremely complicated; political realignments, military conquests, and dynastic replacements occurred frequently. Although increasingly large territorial units and empires were formed periodically, these large polities often were short-lived. In other instances, expansive polities came into contact and conflict with one another, leading to the collapse of at least one of them. The political fluidity that has characterized this part of the world in recent decades may have roots in the distant past.
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 451
Concept
<
Early Writing Systems The economic basis of Mesopotamian written symbols
Early civilizations employed different systems of communication, information storage, and accounting. In the Andes, the quipu was used by Inca bureaucrats to keep accounts; no system of writing as we traditionally conceive of them was developed (see “Huánuco Pampa,” Chapter 9, p. 429). Some of the first writing in China was on animal bones and turtle shells. Scapulas (shoulder blades) and other bones were heated until cracks formed. The patterning of the cracks in relation to the written characters was used ritually by diviners to foretell the future, a practice referred to as scapulimancy (see “An-yang,” p. 475). In the Indus Valley, evidence for early indigenous script is found primarily on soapstone seals that apparently were used to mark ownership. In Egypt, many early texts recorded dynastic and kinship themes. In prehispanic Mesoamerica, writing was frequently used to record dynastic births, marriages, accessions, and deaths; to announce military victories; and to document political events. Calendrical inscriptions often were used by the Maya and other Mesoamerican peoples to track historical themes. In Mesopotamia, the principal function of the world’s first writing system was similar to that described for the Indus Valley; however, the particular conventions of the Mesopotamian writing system and the actual script and symbols employed were markedly different. In 1929, a team of German archaeologists discovered written inscriptions on clay tablets at the site of Uruk. These earliest Mesopotamian texts, dating to the late fourth millennium B.C., revealed an already-developed writing system that included as many as 1500 different symbols and fairly consistent conventions for the presentation of information. By 3000 B.C., this writing
Figure 10.13 A clay envelope, or bulla, approximately 7 cm (2.8 in) in diameter, and examples of the small clay tokens that were placed inside it.
system was in use across southern Mesopotamia. The written symbols were primarily ideographs (abstract signs), although a few of the signs were pictographs that more or less portrayed the represented objects. In the mid-1960s, French excavations at Susa, on the Susiana Plain east of Uruk, discovered hollow clay spheres or envelopes, called bullae, that enclosed modeled clay tokens or geometric forms (Figure 10.13). Dating to the end of the fourth millennium B.C., these bullae are impressed on the outside, thus providing evidence for writing that is approximately as old as the Uruk tablets. They resemble the hollow clay tablets containing clay tokens that were used for accounting as late as 1500 B.C. in parts of Southwest Asia. In 1969, Denise Schmandt-Besserat, of the University of Texas at Austin, initiated a study of the earliest uses of clay in Southwest Asia. Traveling to museums around the world, she recorded collections of early pottery, bricks, and figurines. Yet in addition to those expected objects, she noted that
scapulimancy The ancient practice of
seeking knowledge by reading cracks on bones. Symbols were written on an animal’s scapula (shoulder blade); the bone was heated until a series of cracks formed; then diviners interpreted the pattern of cracking to foretell the future. soapstone A soft stone with a soapy
feel that is easy to carve; often referred to as steatite. bulla (plural bullae) (Latin) A hollow
clay sphere or envelope used to enclose clay tokens in ancient Mesopotamia.
Concept Early Writing Systems
451
_
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 452
Figure 10.14 Examples of clay tokens. Some of the tokens can be equated with early Sumerian ideographs; for example, the bottom left token may refer to metal, the center left to oil, and the center right to sheep.
Recordkeeping enables better recall of past outcomes, promotes reputation formation, and encourages spontaneous coordination of economic decisions. The ultimate effect is that recordkeeping alters an economy’s history and encourages exchange by reducing the risk of loss from transacting with strangers. —Sudipta Basu, John Dickhaut, Gary Hecht, Kristy Towry, and Gregory Waymire (2009)
0
2 cm
0
2 cm
most of the collections (made from Turkey to Pakistan) also included handmodeled clay tokens dating to as early as the ninth millennium B.C. SchmandtBesserat recognized the similarity between these tokens, which generally had been cataloged as toys, gaming pieces, or ritual objects, and the tokens associated with the later bullae and hollow tablets. She also noticed more than 30 formal correspondences between three-dimensional geometric tokens and the two-dimensional ideographic symbols that were inscribed on the early tablets. Based on these correspondences, Schmandt-Besserat reasoned that Neolithic clay tokens, like the later bulla-
_ _
452
Chapter 10 States and Empires in Asia and Africa
enclosed counterparts, may have been used to record economic transactions. The earliest token shapes that could be associated with later written characters referred to quantities of various agricultural products. Other common token forms were linked formally with the two-dimensional symbols for numbers and key commodities, such as cloth, bread, animals, and oil (Figure 10.14). Not surprisingly, the variety of tokens increased through time, generally becoming more elaborate. Overall, more than 200 kinds of tokens have been identified at sites dating to 9000–1500 B.C. If Schmandt-Besserat’s hypotheses are correct, the first communication revolution—the advent of the clay
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
tokens—was associated with the transition to village sedentism, farming, and an increased volume of economic transactions and long-distance trade. Relatively little change in this recording system is evident before the emergence of urbanism and the state during the fourth millennium B.C. Sometime after 3500 B.C., a sizable proportion of the tokens were perforated, as if for stringing. Schmandt-Besserat reasoned that tokens may have been strung together to signify that the objects they represented were part of a single transaction. By the end of the fourth millennium B.C., tokens also began to be placed in bullae. Sealing the tokens within a single bulla could have served to segregate a specific transaction. Seals of the individuals involved were placed on the outside of the bulla, perhaps validating the event. A bulla may have been used as a bill of sale would be used today. The honesty of the deliverer could be checked by matching the goods received with the tokens enclosed in the accompanying bulla. But there was one major drawback to this innovation: Checking the tokens required that a bulla be broken; but to preserve the record, the bulla had to remain intact. The solution was to press the tokens on the exterior of the clay envelope before enclosing the tokens and firing the bulla. In this way, the contents of the load could be checked by the receiver while the validating inscriptions remained intact. In some instances, a finger or a stylus was used to impress the image of the token on the bulla surface. By this time, a system of cylinder seals also had been devised. The seals could be impressed on clay, leaving a mark or design that could be associated with a specific person. By the outset of the third millennium B.C., hollow bullae with tokens inside had generally been replaced by clay tablets. Anyone could check the outside of a bulla without destroying the record. An inscribed bulla or tablet was much easier to make and store than clay tokens. Such economizing measures would have become more important as the development of tribute-collecting institutions (the temple) and urban centers fostered
Page 453
an intensified volume of economic transactions. Thus it is not really surprising that the earliest tablets were convex and made of clay, mirroring the shape and material composition of the spherical bullae. Likewise, it is easier to understand the early reliance on ideographs on the tablets, because the earliest written signs followed the shapes of the geometric tokens. The use of Mesopotamian writing to record economic transactions also seems a logical development from the function of the earlier token system. Mesopotamian writing clearly did not emerge in a vacuum; it evolved from an old and widespread system of communication and accounting. Mesopotamian writing was simplified and made more efficient over time. To simplify writing, Sumerian scribes reduced the number of symbols and substituted wedge-shaped marks for the signs, giving rise to cuneiform, the name assigned to subsequent Mesopotamian scripts. The opposite occurred in Egypt, where the hieroglyphic symbols were made increasingly difficult to execute. In Egypt, writing served fewer purposes for a relatively small segment of the population. The most significant contribution of writing was expediting the flow of information in increasingly large and stratified societies. It facilitated administrative activities and enabled the further growth and centralization of Mesopotamian cities. Writing also may have crystallized and preserved Mesopotamian cultural and bureaucratic traditions so that they outlived the hegemony of single rulers or dominant city-states. Yet an alphabetical system was not invented in Mesopotamia, and in 1900 B.C., Sumerian written language still contained 600–700 unique elements, with an organizational structure somewhat analogous to traditional Chinese. The first truly alphabetical written languages developed toward the end of the second millennium B.C. By 1000 B.C., the Greeks had adapted the Syrian alphabet to their own language and reduced the number of written signs to 25. The Greek alphabet then became the foundation for all contemporary European language systems.
What was the nature of early writing in Mesopotamia, and how did it develop?
y
cuneiform A writing system of ancient Mesopotamia involving a series of wedge-shaped marks to convey a message or text.
Concept Early Writing Systems
453
_
pri31057_ch10.pp436-505.qxd
10/15/09
11:06 AM
Page 454
Site
:
Harappa and Mohenjo-daro Urbanism and the rise of civilization in the Indus Valley The broad, fertile floodplains of the Indus and Ghaggar-Hakra rivers (the latter is now dry) and their tributaries, in what is now Pakistan, were the principal focus of the Indus Valley civilization, 2600–1900 B.C. Covering about 680,000 sq km (260,000 sq mi, roughly the size of Texas), this area is bordered by the Baluchistan Hills to the west, the Arabian Sea to the south, the Great Indian Desert to the east, and the majestic Himalayan Mountains to the north (Figure 10.15).
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on Indus Valley civilization, see the Internet exercises on your online learning center.
Figure 10.15 The Indus area, with sites mentioned in the text.
M
N
HI M OU ALA N T YA A IN
S
AFGHANISTAN
Harappa PAKISTAN PUNJAB us
Gh
d In
Mohenjo-daro BALUCHISTAN
R
.
Mehrgarh
ag
ga
R. kra a r-H
G R E AT I N D I A N D E S E RT
INDIA Balakot
Arabian Sea G U J A R AT Lothal
_ _
454
Chapter 10 States and Empires in Asia and Africa
The earliest known sites in the riverine heartland of the Indus civilization date to the late fifth and early fourth millennia B.C., postdating earlier occupations such as Mehrgarh, 200 km (125 mi) to the west. These settlements, referred to as Early Harappan, were scattered across the plains in major agricultural areas or along important trade routes. Many of them exhibit artifacts and organizational features directly antecedent to the later sites, suggesting that the Indus Valley civilization had a long, local path of development. The Indus development does not appear to have been a simple consequence of stimuli from the ancient civilizations of Mesopotamia, a view held by previous generations of scholars. The early Indus Valley settlements consisted of small, contiguous, rectangular mud-brick houses, some of which contained multiple rooms. The size of settlements varied, and a few included monumental construction. Some sites had massive mud-brick walls and neighborhoods laid out with north-south and east-west streets. Plow-based agriculture was practiced. Cattle, sheep, and goats were kept, but hunting and fishing remained important subsistence activities. Craft technologies associated with later Indus civilization developed to a high degree at these pre-Indus settlements. Rings, bangles, beads, pins, axes, and celts were manufactured from copper and bronze. Fine stones were ground and polished into beads. Using kilns and the potter’s wheel, craftspeople produced a variety of vessel forms, some of which were elaborate, such as serving dishes on stands. Much of the pottery was finely painted. Other important crafted items included terracotta figurines. By the late fourth millennium B . C ., potter ’s marks were
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
present, and seals were inscribed with various geometric symbols. Indus Valley civilization, also called the Harappan tradition, was first identified by Sir John Marshall in 1921 at the site of Harappa (ha-RAP-ah) in the Punjab highlands in the upper Indus Valley. Although few systematic settlement pattern studies have been undertaken, 1500 Harappan sites have been reported. Few villages have been excavated, but most appear to be 1–5 ha (2.5–12.5 acres) in size and are located near rivers or streams. There are at least four large urban centers, the best known of which are Harappa and Mohenjo-daro (mo-HENGE-o-DAHro), roughly 500 km (310 mi) to the south in the lower Indus plain. The two sites, both of which have been the focus of major archaeological field studies, are surprisingly similar. Both towns are large, covering approximately 150–250 ha (370–620 acres), and contained populations of roughly 40,000– 80,000 people. Both Mohenjo-daro and Harappa were built with massive mudbrick walls and platforms that raised the towns above the surrounding floodplains (Figure 10.16). Mohenjo-daro was rebuilt at least nine times. Harappa and Mohenjo-daro consisted of several mounded sectors. Massive foundations of eroded mudbrick walls and traces of large brick gateways have been noted around the edges of these mounds. Both Harappa and Mohenjo-daro have a high rectangular mound on the west and other large mounds to the north, south, and east. Some of the most important public buildings associated with the Harappan tradition are located on the western, or tallest, mound at Mohenjodaro. The major structures included a possible “granary,” a “great bath,” and a great hall (“college”) almost 730 sq m (8100 sq ft) in size. Some scholars have argued that the granary (1000 sq m, or 11,000 sq ft, equivalent to an Olympicsize swimming pool) was erected over brick supports so that air could circulate under the stored grain. Others suggest that it was simply a public building with multiple rooms or the
Page 455
residence of a powerful merchant family. The Great Bath—12 ⫻ 7 m (39 ⫻ 23 ft) and 3 m (10 ft) deep, which may have been used for ceremonial ritual bathing—included eight small private bathrooms or changing cubicles (Figure 10.17). The bath itself was fed by a well, and its brickwork was sealed with waterproof bitumen or tar. The building was probably the first large water tank in the world. Mohenjo-daro’s other sectors were divided into blocks by streets, the broadest of which were about 10 m (33 ft) wide (Figure 10.18). Hundreds of houses lined the streets and alleys, some of which were paved with stone. Some of these structures had two stories and were made of baked mud
Figure 10.16 The ruins of ancient Mohenjo-daro, a well-preserved center in the Indus Valley of modern Pakistan. Much of the site was built on a massive mud-brick platform, which raised the settlement over the wet alluvial plain below. The balloon above the site is used by archaeologists to view and photograph the ruins from the air.
_ Site Harappa and Mohenjo-daro
455
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 456
Drain
Tank
0
10 m
Figure 10.17 The plan of the Great Bath at Mohenjo-daro.
_ _
intaglio An engraving in stone or other hard material that is depressed below the surface; an impression of the design produces an image in relief.
456
bricks. (At smaller settlements, houses generally were built of sun-dried mud bricks.) At Mohenjo-daro, more spacious dwellings, perhaps for highstatus individuals and merchants, were laid out around central courtyards. These residences had private bathing areas and toilets connected to a central drainage system built partially underground. A striking feature of Harappan society is the extent of standardization, including a system of weights and measures. Precisely shaped pieces of chert or agate were used as counterweights in balances (Figure 10.19). Construction bricks had standard dimensions. Ceramic forms and ornamentation were remarkably similar at sites throughout the Indus system, although there was some regional variation and change over time.
Chapter 10 States and Empires in Asia and Africa
A considerable degree of occupational specialization characterized Indus society, and one’s profession was probably an important factor in social differentiation. At the major urban centers, there were designated living and working quarters for beadmakers, coppersmiths, and weavers. Certain smaller sites were devoted almost entirely to a specific industry, craft, or trade, including beadmaking, shellworking, ceramic production, and coppersmithing. Metallurgy was well developed, and copper and bronze were used for a variety of tools and weapons. The availability of copper, lead, and silver within or close to Indus territory contributed to a greater use of metal tools than was evident in Mesopotamia. There were significant differences between the Indus Valley and Mesopotamian civilizations. Although the Indus civilization covered a larger geographic area—650,000 sq km (260,000 sq mi)—it had a smaller number of major centers. Mesopotamia was composed of many city-states. The similarities between Harappa and Mohenjo-daro suggest that Indus centers were closely linked economically and culturally. The Indus civilization may have had a more equitable distribution of wealth than was the case in other early Eurasian societies. Exotic stones and metal were not restricted to large sites or to clearly elite contexts; considerable quantities of wealth have been recovered at even modest settlements. Indus material culture was simple compared with that of Mesopotamia; little representational or lavish art was constructed on a massive scale. Instead, one finds figurines, small sculptures, carvings on bone and ivory, decorated pottery, and intaglio figures on seals. Indus art is often in miniature. The Indus elite engaged in fewer lavish public displays; they built no rich tombs, elaborate palaces, or fancy temples. But some mud-brick-lined tombs yielded more grave offerings (pottery, bronze mirrors, and a few beads) than the average burial. Nothing approaching the royal graves at Ur or the Egyptian tombs has yet been
pri31057_ch10.pp436-505.qxd
Figure 10.18
10/1/09
8:21 AM
Page 457
The plan of streets and houses in one sector of Mohenjo-daro.
Figure 10.19 A collection of stone cubical weights from Harappa.
0
5 cm
_ Site Harappa and Mohenjo-daro
457
pri31057_ch10.pp436-505.qxd
10/1/09
Figure 10.20 A steatite figure of a ruler from Mohenjo-daro, standing 19 cm (7.4 in) high; such depictions of rulers are rare in Harappan civilization.
chlorite A kind of green stone that
resembles mica. serpentine A stone of dull green
_ _
color that often has a mottled appearance.
458
8:21 AM
Page 458
found in the Indus region. In fact, whatever rulers and elite the Harappans had remain anonymous. Individual conquests and accomplishments are not enumerated, and few portraits have been found. A steatite (soapstone) figure from Mohenjo-daro depicting a bearded man in an embroidered robe is a rare exception (Figure 10.20). Indus settlements were closer to natural resources than Sumerian sites were. The large Harappan centers were connected with outlying rural communities and resource areas through complex trade networks. Both maritime and overland routes were used. Wheeled
Chapter 10 States and Empires in Asia and Africa
carts were drawn along regular caravan routes, and deep-sea vessels traveled the Persian Gulf and the Gulf of Oman. Long-distance exchanges moved seals, carnelian beads, and other miscellaneous items from the Indus region to the Persian Gulf, northern and southern Mesopotamia, Iran, and Afghanistan. Imports included lapis lazuli from Afghanistan, conch shells from Gujarat (western India), turquoise from northeastern Iran, carved chlorite bowls from the Iranian Plateau, and serpentine from central Asia. Few Mesopotamian items have been found in the Indus area. A possible explanation is that Mesopotamia exported mostly perishables or specific raw materials, such as barley, fruits, oil, and textiles. A system of writing that was very different from the early Mesopotamian script developed in the Indus Valley. Over 4000 seals with Indus script have been found (Figure 10.21). Most inscriptions are short; no known inscription is longer than 21 signs, and the average text is only 5 or 6 signs. The lack of long texts has added to the difficulty of interpreting the Indus script. Over 400 symbols have been identified, yet none has been definitively deciphered. Inscriptions, including both writing and pictures, are found on small copper tablets and potsherds, but most are found on square seals of soapstone. Most seals had holes that allowed them to be strung and worn around the neck. The script seems to identify the owner of the seal or the official status of the bearer. Many animals are depicted on the seals, including elephants, water buffalo, rhinoceroses, tigers, crocodiles, antelopes, bulls, and goats. Walter Fairservis, Jr., suggested that these animals may be totems or symbols representing specific kin groups; some seals depict processions with animal effigies being carried as standards. Each seal may thus identify its owner with a social group, which might help explain why similar scenes are repeated on multiple seals. One common theme is a figure seated in the yogic posture with
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 459
Figure 10.21 A Harappan seal with a bull in profile, a common motif on the seals; this medium-size square seal is 2.5 cm (1 in) high.
heels pressed under the groin. Surrounded by various animals, this individual wears a water buffalo–horned headdress. A second common scene is a pipal tree with various anthropomorphic and human figures. Some archaeologists have suggested that the figure with the headdress may be an early form of a deity that later came to be worshipped as Shiva, Lord of the Beasts, in Hinduism. The Indus civilization began to decline around 1900 B.C. The major Harappan centers were greatly weakened, and the center of power shifted from the Indus to the Ganges River Valley to the east, where, after 600 B.C., large cities were built and state-level organizations formed. Archaeologists now think that the changes in the early part of the second millennium B.C. were not a complete collapse or population replacement, but rather the beginning of an episode of decentralization. Many elements of the earlier Harappan civilization were retained in these new settlements. The decline of the Indus civilization has been linked to the
drying up of the ancient Ghaggar-Hakra River and the breakdown of the Indus system of exchange. Traditional views that proposed an Aryan invasion from the northwest are no longer supported. The persistence of many aspects and traditions of the Indus civilization into more recent times is startling. Ceremonial bathing, ritual burning, specific body positions (such as the yogic position) on seals, the important symbolic roles of bulls and elephants, decorative arrangements of multiple bangles and necklaces (evident from graves and realistic figurines), and certain distinctive headgear—all are important attributes of ancient Harappan society that remain at the heart of contemporary Hinduism. The standard Harappan unit of weight, equivalent to 14 g (0.5 oz), continued in use at South Asian bazaars and markets into the nineteenth century. Although ancient Harappa and Mohenjo-daro lie in ruins today, the civilization of which they were a part has left an extremely important legacy.
Compare and contrast the Indus Valley and Mesopotamian civilizations. What are some of the major differences between these two early civilizations?
y
pipal tree A species of fig tree on the
South Asian subcontinent that has had sacred significance for many cultures and religions throughout the region for thousands of years.
_ Site Harappa and Mohenjo-daro
459
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 460
Concept
<
Economic Specialization Shellworking in the Indus civilization
Craft goods were extraordinarily important in the production and maintenance of ancient chiefdoms and states. In addition to basic domestic functions, they were used in almost all social, political, and ritual activities. Understanding the context and organization of their production is integral to a full understanding of daily life, political economy, and the role of material objects in social and political relations. —Cathy L. Costin (2004)
_ _
division of labor The differentiation of economic roles in a specific sociopolitical context.
460
With population growth, more densely packed communities, and increasing political complexity, household selfsufficiency becomes difficult, if not impossible, to sustain. Gaining access to all the resources that a family requires becomes difficult. As more tasks become specialized (a division of labor), people exchange what they produce for other goods and services through markets and trade networks. In some societies, the taxing or control of markets or exchange networks is an important source of power. Today when we think of craft production or the specialized manufacture of goods for exchange, we often presume that such activities occur in nondomestic workshops or factories. Yet in many regions of the world, preindustrial economic specialization was largely situated in residential contexts or houses. Over the past few decades, archaeological investigations have devoted increasing attention to identifying the material indicators of craft manufacture, as well as the context of the locations where that production was situated and the volumes at which such craft products were made. Exchange and craft production appear to have been key features of Indus Valley society. Complex internal trade networks connected the major urban centers of the Indus civilization with rural agricultural and resource areas. Evidence from Mohenjo-daro suggests that the large Indus cities included craft areas that served as the living and working quarters for specialists. Some crafts, such as the working of shell, stone, pottery, and metal, may have developed into hereditary occupations. Shellworking, an important Indus craft, was undertaken by specialists. The earliest use of shell was limited to simple ornaments that were made by perforating natural shells. Later, dur-
Chapter 10 States and Empires in Asia and Africa
ing the time of the Indus civilization, shell use increased to include a variety of such decorative, utilitarian, and ritual objects as ornaments (bangles, rings, beads, pendants, and large perforated disks), utensils (ladles), inlay pieces, and other special objects. According to J. Mark Kenoyer, of the University of Wisconsin–Madison, each of the workshops at Mohenjo-daro specialized in producing different shell items. For example, one area apparently produced mostly inlay pieces. Shell workshops similar to those at Mohenjo-daro were present at Harappa and other urban centers. But at Harappa, there was less variety in shell species and fewer shell artifacts in general, because of its location further inland. At the site of Lothal, on the coast of the Arabian Sea, shell workshops also produced a variety of shell objects. Another major shell site, at Balakot, on the coast near Karachi, Pakistan, specialized in shell bangles, beads, and smaller objects. The site has workshop areas with stone grinders and hammers, bangles in various stages of manufacture, and unworked shell. One type of shell was cut with a specialized bronze saw. Metal tools were expensive and, at most sites, only craftworkers who were supported or controlled by more affluent individuals had access to metal tools. Most of the bangles were made by an alternative chipping and grinding process that used stone tools (Figure 10.22). Regardless of the method of manufacture, the resulting bangles at Balakot were almost identical. Even though Indus sites specialized in different types of finished products, a single standardized manufacturing technology and certain decorative conventions often were employed across the region. Although certain shell items were purely decorative, the function of other
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 461
Figure 10.22 The manufacturing process from large whole shell to finished bangles. The process begins with the preliminary chipping of the shell and the removal of the internal columella (a–c) and continues with sawing the body of the shell into thin circles (d–g), finishing the edges of the shell blanks (h–j), and incising the final bangle (k).
(b) (a)
(d)
(e)
(c)
(f)
(i)
(g)
(h)
(j)
Indus shell artifacts remains a mystery. Recent excavations at the cemetery area in Harappa have found many adult women with shell bangles on their left arms. These arm bracelets may have been a symbol of ethnic identity or a signifier of a specific marriage status. Shell bangles are still used for various social and ritual functions across the South Asian subcontinent. Through historical accounts, the antiquity of
(k)
finely crafted shell objects (and their ritual functions) can be traced back to 600 B.C. It seems reasonable to deduce that some of these social and ritual uses may have their ultimate roots in the practices of the Indus civilization.
_ Concept Economic Specialization
461
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 462
Site
:
Hierakonpolis The emergence of the Egyptian civilization In spite of the general popularity of Egyptian archaeology, we know surprisingly little about early Egypt, particularly before the rise of early states and the first written documents. What we do know about Egypt suggests that this civilization along the Nile River was different in its long-term history from the ancient civilizations that developed in other parts of the world.
Figure 10.23 The Nile Valley in Egypt, with sites of Upper and Lower Egypt mentioned in the text.
M e d i t e rran ean S ea
THE
NILE DELTA
LOWER EGYPT
Merimde Giza
Memphis
N i l e River
El Gerza
ARABIAN DESERT
UPPER EGYPT
El Badari
LIBYAN
Red Sea
DESERT
This Abydos El Amra Valley of Kings
Naqada Thebes Hierakonpolis
0
160 km First Cataract
Aswan
_ _
462
Chapter 10 States and Empires in Asia and Africa
For example, Egypt was relatively centralized for almost 2500 years, with only one major episode of political fragmentation, 2200–2000 B.C. Egyptian civilization centered on the Nile Valley, a long oasis surrounded by desert (Figure 10.23). The Nile flows to the north and the Mediterranean Sea, more than 6400 km (4000 mi) from its source in the swamps and lakes of equatorial Africa. Its final 1300 km (800 mi) cut through Egypt before fanning out into an enormous delta. In Lower Egypt, an area of rich cultivable floodplains to the north, the river valley is up to 20 km (12.5 mi) wide. The south, known as Upper Egypt, has less alluvial land (only several kilometers wide), forming a narrow strip surrounded by jagged rock escarpments. Before the construction of the Aswan Dam, near the First Cataract (rapids), in the 1960s, annual flooding along the Nile was common, and tons of rich soil were deposited. As a result, the floodplain of the Nile is extremely fertile. Because of sparse rainfall today, irrigation is necessary for farming, but annual temperatures are ideal for the cultivation of a wide range of crops. Staple crops are legumes, barley, onions, cucumbers, melons, and figs. Other plants include rushes and reeds used for making baskets, flax for linens, and papyrus for cordage and paper. As in the past, today the area also supports sheep, goats, pigs, cattle, ducks, geese, fish, turtles, crocodiles, hippopotamuses, and other game animals. Although the surrounding deserts provide very little in the way of food, they are rich in building stone and minerals, including copper, gold, and silver. The floods along the Nile were more predictable and easier to control
pri31057_ch10.pp436-505.qxd
10/15/09
11:06 AM
than those of the Tigris and Euphrates rivers in Mesopotamia. The Egyptians could very easily modify natural basins on the floor of the Nile Valley to retain floodwater for their crops. Higher levees along the river provided dry locations for settlements. No definite sedentary villages have been found in the Nile Valley before the sixth millennium B.C., when nomadic cattle herders and farmers began to settle in the area. Before 6000 B.C., the Nile Valley was occupied by groups of huntergatherers who followed an annual round in which they hunted along the margins of the desert for wild cattle, gazelles, and birds during part of the year and took fowl and fish at other times. The presence of grinding stones indicates that wild grains also were an important food item. Soon after 5000 B.C., food production was established in the Nile Valley. The early farming settlements raised Southwest Asian domesticates: wheat, barley, sheep, and goats. Most of what is known about predynastic Egypt comes from the south, where there has been less deposition to bury sites. Yet it is possible that sedentism actually occurred earlier in the more fertile north. One floodplain settlement near the Nile Delta, Merimde, dating to as early as 4900 B.C., consisted of a cluster of semisubterranean, oval houses with roofs of sticks and mud. The inhabitants used stone axes, knives, and flint arrowheads. Grains were stored in ceramic jars, baskets, and pits. Circular, clay-lined threshing floors also have been reported. The earliest occupations in the south (roughly contemporary with those in the north) are called Badarian, after the best-known settlement of El Badari in Upper Egypt. Badarian settlements consisted of clusters of skin tents or small huts. Many of the dead were buried carefully in oval or rectangular pits roofed over with sticks or mats. Utensils and food were commonly placed in the burials. Other grave offerings included rectangular stone palettes (or tablets), ivory spoons, and small
Page 463
ivory or stone vases, all of which have been associated with the preparation and use of green face paint. These items are all common components of the predynastic burial assemblages. The Badarian focus on grave features and the accompanying goods may have been at the root of the later Egyptian emphasis on burial custom. In the south, materials associated with the Amratian tradition (3800– 3500 B.C.) are frequently found directly above Badarian levels. The name Amratian was derived from the site of El Amra near Abydos in Upper Egypt. Although the Amratian materials seem to reflect direct continuity with the earlier Badarian, the latter period is characterized by the appearance of more developed craft industries (particularly pottery, alabaster, and basalt) and a larger number and more widespread distribution of settlements. Copperworking, which may have been initiated in the Badarian, also gained in importance. Amratian metalworkers used copper to make pins, flat axes, awls, and daggers. Scholars do not agree on the origins of Egyptian metallurgy. Many argue that it was introduced from Southwest Asia, whereas others, pointing to earlier metalworking in Upper than in Lower Egypt, suggest that it may have been an indigenous accomplishment. During the Amratian period, at least three important centers emerged in southern Egypt: Naqada, This (near Abydos), and Hierakonpolis (HIGHra-KON-po-lis). The best known of these, Hierakonpolis, had an estimated population of several thousand people. The Greek name for this settlement, which means “city of the hawk,” comes from the falcon-headed Egyptian god Horus of the city Nekhen, the Egyptian name for this ancient center. Most of the site’s inhabitants lived in rectangular, semisubterranean houses of mud bricks and thatch. The more important artisans and traders lived in larger houses in separate compounds. Near the site was a large cemetery with part of the burial area reserved for
papyrus A tall marsh plant, or reed, of the Nile Valley that the ancient Egyptians cut into strips and pressed into a kind of paper to write on.
_ Site Hierakonpolis
463
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Figure 10.24 The ceremonial macehead found at Hierakonpolis, shown reconstructed here. The scorpion in front of the face of the principal individual, who wears the white crown of Upper Egypt, identifies him as King Scorpion, the predecessor of Egypt’s first pharaoh.
_ _
porphyry An igneous rock with visible quartz or feldspar crystals embedded in a finer-grained base.
464
Page 464
elaborate tombs. Although not as spectacular as later ones in Egypt, these tombs are impressive. The early tombs were constructed by cutting rectangular holes into the terrace of a dry streambed. The size and contents of the tombs varied, perhaps reflecting status and prestige. One grave, measuring 2.4 ⫻ 1.5 m (8 ⫻ 5 ft) and 1.8 m (6 ft) deep, contained baskets, fine pottery, rope, flint arrowheads, and wooden arrow shafts. Unfortunately, the grave was looted in later times, and the most valuable goods may have been stolen. A second grave contained scraps of papyrus paper and a disk-shaped macehead made of polished green-and-white porphyry. Maceheads, which were mounted on wooden staffs called maces, were recognized in Egypt as a sign of authority (Figure 10.24). The presence of the macehead at Hierakonpolis indicates that the process of political development already may have been well under way. Similar polished, hard-stone mace-
Chapter 10 States and Empires in Asia and Africa
heads have been found at several other contemporaneous sites, possibly indicating the presence of several small, competing political units. Hierakonpolis also was the center of a very large pottery industry. At least 15 Amratian kilns have been identified, the largest of which covered over 1000 sq m (0.25 acre). Two distinct types of pottery were made: a coarse ware for everyday household or industrial use and a fine, untempered ware for grave offerings. The kilns appear to have been part of one well-organized complex, in that each kiln was used for the firing of a different kind of pottery. The large kilns must have produced far more than was needed locally. Michael Hoffman postulated that the people in charge of pottery production acquired considerable economic power. During the subsequent Gerzean period (3500–3100 B.C.), named after El Gerza in the north, craft activities, including pottery production, metallurgy, and the manufacture of stone bowls from very hard materials such as diorite and basalt, appear to have been carried out on an even larger scale. The use of copper artifacts, which were now cast as well as hammered, increased. Gold also was worked at this time, and some luxury items were wrapped in gold foil. Trade with Southwest Asia intensified in volume. Some of the pottery of this period was painted in Southwest Asian style, in dark red on a buff background. Foreign motifs also were incorporated into the local decorative tradition. Gerzean remains have been found in Upper and Lower Egypt, possibly indicating greater integration between the two areas. Yet important differences in ceramic styles and burial customs continued to distinguish Upper and Lower Egypt. On carved palettes and other art, key individuals are depicted wearing distinctive headgear or crowns. These crowns vary from one area of Egypt to another (Figure 10.25), suggesting that several different polities may have developed along the Nile. During Gerzean times, social and economic inequalities, evident from
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
tomb size, grave design, and burial inclusions, increased markedly. One macehead found at Hierakonpolis, with a scorpion on it, has been identified as belonging to King Scorpion, the predecessor of Egypt’s first pharaoh (see Figure 10.24). Hierakonpolis, whose location shifted slightly closer to the Nile, was one of the largest known settlements, with 5 ha (12.5 acres) of occupation, including large nonresidential structures such as palaces and temples. Written records and stone monuments indicate that the incidence of warfare increased at that time, with local kings trying to gain control over adjacent kingdoms. Among the competing kings, it is thought to have been Narmer from Hierakonpolis who finally succeeded in unifying Egypt into one kingdom around 3100 B . C . As Egypt’s first pharaoh, Narmer founded a dynasty (3100–2890 B.C.) and a political structure that lasted for nearly 3000 years. The unification of Upper and Lower Egypt was recorded for posterity on a 64-cm (25-in) carved stone palette discovered by English archaeologists in 1898 at Hierakonpolis (Figure 10.26). One side of the tablet depicts Narmer, the leader of Upper Egypt, wearing the white crown of Upper Egypt and holding his symbolic mace. The other side shows him wearing the red crown of Lower Egypt. Both sides include scenes of the king in battle. The Egyptian state was far larger and more complex than any city-state in Mesopotamia. The nature of rule also was much different in Egypt than in Mesopotamia. In Egypt, the royal court centralized power and wealth, and this consolidation is evident in the concentration of resources in the mortuary complexes of the kings. The relative stability of Egyptian rule diminished some of the insecurity that in Mesopotamia led to the construction of great walled complexes around nucleated urban centers. In Egypt, continuity with earlier occupations is apparent in such contexts as stoneworking, death rituals, and the use of the macehead as a symbol of rule.
Page 465
The unification of Egypt was closely timed with the earliest hieroglyphic writing. Although writing first developed in Mesopotamia, the Egyptians devised their own script, which is very different from writing in Southwest Asia. Egyptian hieroglyphs consist of both pictographic and phonetic (vocal sounds) elements, written on papyrus, carved on stone as part of public buildings, or painted on wood or clay (Figure 10.27). Egyptian writing was more concerned with rule and kinship than with economic transactions. The deeds and accomplishments of leaders were recorded, and this may have had a role in the consolidation of power in the hands of a limited few. The widespread adoption of irrigation agriculture also coincided with the unification of Egypt. Ancient Egyptian irrigation practices were very simple, such as modifying natural basins to serve as reservoirs for floodwater. Yet the large surpluses that these techniques permitted clearly were necessary to support the opulent lifestyles and mortuary rituals of the emergent pharaohs.
Figure 10.25 The Egyptian crowns: the white crown of Upper Egypt (left), the red crown of Lower Egypt (center), and the double crown of unified Egypt (right).
_ Site Hierakonpolis
465
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 466
Figure 10.26 Narmer’s carved stone palette. One side (left) depicts Egypt’s first pharaoh wearing the white crown of Upper Egypt and posturing over a defeated enemy in the presence of the falcon god Horus. On the other side (right), he wears the red crown of Lower Egypt and marches in a victory procession to view decapitated prisoners.
Scarab beetle (kheper) means “to become” or “to evolve”
The ankh, a cross with a looped top, means “life” or “to live”
Figure 10.27 Some frequently used Egyptian hieroglyphs
_ _
466
Chapter 10 States and Empires in Asia and Africa
Scepter (was), a forked staff topped with an animal’s head, refers to “power”
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 467
Concept
<
The Cemetery at Hierakonpolis Steps toward the unification of Egypt
The earliest tombs in Egypt were small wooden graves in which the dead were buried with utensils and food to sustain them in the afterworld. In late predynastic Egypt (the fourth millennium B.C.), more elaborate tombs were constructed for the burial of elite members of society. These graves contained a variety of goods, including beautiful pottery, baskets, braided leather rope, painted reed arrow shafts, flint arrowheads, pieces of papyrus, and maceheads. The architectural styles of tomb construction differed in Lower Egypt and Upper Egypt (Figure 10.28). In the royal necropolis at Hierakonpolis, these differences were utilized symbolically by Narmer or his immediate predecessors. Constructed at the end of the fourth millennium B.C., the cemetery was arranged, according to Michael Hoffman, who excavated at the site, to represent the union of Upper and Lower Egypt, and perhaps to legitimize the military conquest of Lower Egypt by the Upper Egyptian rulers. Tombs in the style of Lower Egypt were constructed at the downstream end of a wadi, which represented the downstream end of the Nile, or Lower Egypt. These tombs were lined with mud bricks, over which were placed large, painted wood and reed structures. Low reed fences surrounded the tombs. The burials occurred in groups of three, four, and five, suggesting that rulers were buried with other family members or members of their court. A fragmented wooden bed was found in one tomb. It was finely carved, with two legs crafted to resemble a bull’s legs—a forerunner of the furniture in King Tutankhamen’s tomb. At the opposite end of the necropolis, or upstream with respect to the wadi, was a stone tomb in the architectural style of Upper Egypt. This stone feature is surrounded by apparently similar
tombs that have yet to be excavated. Cut into subsoil bedrock, the excavated tomb was a long, narrow trench with an L-shaped hole cut into the middle of the floor. There was no superstructure associated with it. Surrounding this stone tomb were animal burials, including the remains of hippopotamuses, elephants, crocodiles, baboons, cattle, goats, sheep, and dogs. Some of the animals were mummified and probably had civicceremonial significance. In Dynastic times, the Egyptians used various animals as godlike symbols of the different nomes (provinces) that constituted the Egyptian state. Some of these same animal symbols are portrayed atop standards on Narmer’s stone palette. In accordance with the rest of the cemetery, it seems possible that the animal burials also followed a symbolic orientation. For example, the easternmost tomb had six baboons, animals that ancient Egyptians associated with the rising sun. Hoffman proposed that this desert necropolis presages key attributes of the Egyptian state. These features include the use of the royal death cult and an associated cemetery area as national symbols, the division of the state into the symbolic halves of Upper and Lower Egypt, and the incorporation of conquered polities and peoples through ritual as well as military means. In contrast to the neighboring polities of Southwest Asia, political integration in Dynastic Egypt was tied less directly to specific fortified towns and was linked more closely to elite lineages and cemetery and temple complexes. The roots of that civic-ceremonial organization appear to have extended to Hierakonpolis.
Figure 10.28 The two tomb styles of Lower Egypt (bottom) and Upper Egypt (top).
necropolis (Greek) Cemetery. wadi (Arabic) A dry streambed. nome A geographic province incorporated within the ancient Egyptian state.
Concept The Cemetery at Hierakonpolis
467
_
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 468
Site
:
Giza and Dynastic Egypt Pyramids and pharaohs
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the mummification and funerary practices of the ancient Egyptians, see the Internet exercises on your online learning center.
Figure 10.29 Zoser’s stepped pyramid at Saqqarah was an important first step in the construction of monumental pyramids in Egypt.
By 3100 B.C., Upper and Lower Egypt unified into one state when Narmer, Egypt’s first pharaoh, conquered the northern delta. Narmer moved the capital from Hierakonpolis, in Upper Egypt, to Memphis, at the junction of Upper and Lower Egypt, where the Nile Valley spreads out into the broad delta. Memphis remained the political center of Egypt for 1500 years. Although the Nile Valley around Memphis was not a particularly rich agricultural area, the city was strategically positioned for riverine communication between southern and northern Egypt. The symbol of the pharaoh was the double crown of Upper and Lower Egypt (see Figure 10.25). A series of pharaohs during the first (3100–2890 B . C .) and second (2890–2686 B.C.) dynasties ruled under the double crown, but consolidation of the two disparate
_ _
468
Chapter 10 States and Empires in Asia and Africa
regions was not a steady or easy process. The Egyptian kings adopted the strategy of establishing outposts, temples, and shrines throughout their domains, maintaining integration and preventing the monopoly of functions or power in any single place or capital. Intermarriage of the elite from north and south may have been used to solidify the unification. The Egyptian population was less urban than Southwest Asian societies, which generally had larger centers. Apart from a few large sites, most of the Egyptian population continued to live in unwalled, largely self-sufficient villages. This may account for the preoccupation of the Egyptian elite with rural lifeways. Despite this preoccupation, a massive, hereditary bureaucracy developed, devoting official energy to tax collection, harvest yields, and the administration of irrigation. Each nome, administered by a local governor, was under overall central control. During the first dynasty, the Egyptian kings also supported an increasing number of craft specialists. Trade links were extended to what is now Sudan (Nubia) and Libya. With the elaboration of the court during the early dynasties, the demand for subSaharan products such as ivory and ebony intensified, thereby heightening interest in areas to the south. Between the third dynasty in 2686 B . C . and the Persian conquest in 525 B.C., Egypt was ruled by no fewer than 23 dynasties. The third through sixth dynasties constitute what is known as the Old Kingdom (2686–2181 B.C.), a time of despotic pharaohs and grandiose pyramid construction. The largest Egyptian pyramids were constructed by these early dynasties. The first pyramid was a stepped stone structure constructed by the pharaoh Zoser as the centerpiece of his funeral complex (Figure 10.29). The step pyramids were soon
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
followed by the more familiar pyramids with smooth faces. Like all royal tombs until 1000 B.C., the pyramids were constructed in the desert on the west side of the Nile River and were surrounded by the tombs of contemporary officials. Although few written records have survived from the Old Kingdom, sketchy accounts provide a perspective largely unavailable for earlier periods. Scribes were an important part of the government. Special schools trained writers for careers in the palace and the treasury. Contemporary documents on papyrus show that Egyptians were skilled in architecture, surgery, accounting, geometry, and astronomy. But most of the population, as in other early civilizations, consisted of illiterate peasant farmers who maintained the Egyptian agricultural base. One major site of the Old Kingdom is Giza, located near modern Cairo (Figure 10.30). Just a short distance from the ancient capital city of Memphis, the Giza Plateau is where King Khufu (known as Cheops in Greek), of the fourth dynasty, built his massive pyramid. Called the Pyramid of Khufu, or the Great Pyramid, this monumental edifice required precise planning and complex engineering and serves as a clear reflection of state power and labor control. The pyramid is 150 m high (500 ft, almost as tall as the Washington Monument), covers 5.3 ha (13.1 acres), and contains a series of internal passages and chambers. Khufu’s pyramid was constructed of 2.3 million stone blocks with an average weight of 2275 kg (2.5 tons). Its construction involved roughly 13.4 million man-days of labor and incorporated materials from areas as far away as Lebanon, Sinai, Aswan, and Nubia. Surrounding the Pyramid of Khufu were the much smaller pyramids of three queens and the rectangular tombs of Khufu’s closest royal relatives and high officials (Figure 10.31). Nearby was the king’s palace complex and villages for the administrators and workmen. Two other large pyramids were constructed by his successors Khafre
Page 469
(known as Chephren in Greek) and Menkaure (Mycerinus in Greek) (Figure 10.32). Khafre’s pyramid is almost as large as Khufu’s, but its internal structure is much simpler, with a single tomb chamber at the base of the structure. One of the most famous ancient Egyptian structures, the king-headed lion, or Great Sphinx, bears Khafre’s features. It was constructed to guard the sacred realm of the dead kings, whose power was believed to have continued to influence the universe in which their successors ruled. Although of the same design as the large pyramids, Menkaure’s pyramid is much smaller, and its construction marks the end of the era of massive pyramid construction. Those built during the following dynasty were relatively small and poorly constructed. During the early dynasties, pharaohs were central to the Egyptian state. They were considered divine, sometimes referred to as “the good god.” These paramount rulers controlled economic exchange, served at the top of a great bureaucracy, and acted as the heads of state
Figure 10.30 The enormous pyramid complex at Giza, near modern Cairo, built during Old Kingdom Egypt (2686–2181 B.C.). The pyramids were constructed over a long period of time by a succession of pharaohs, each of whom had his own funerary monument.
_ Site Giza and Dynastic Egypt
469
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 470
Figure 10.31 The plan of Giza, showing the three large pyramids of the pharaohs, the smaller pyramids of the queens, and other smaller tombs.
Avenue of the Pyramids
Pyramid of Khufu
Western tomb field
Later Egyptian records mention at least seventeen kinds of beer, some of them referred to in poetic terms that sound, to modern ears, almost like advertising slogans: Different beers were known as “the beautiful and good,” “the heavenly,” “the joy-bringer,” “the addition to the meal,” “the plentiful,” “the fermented.” Beers used in religious ceremonies also had special names. —Tom Standage (2005)
Boat pits
Eastern tomb field
Boat pits
Builders’ quarters?
Pyramid of Khafre Mortuary temple Rockcut tombs
Tombs
Rock-cut tombs
Pyramids of queens
Caus
Great Sphinx
eway
Sphinx temple Central field of rock-cut tombs
Valley temple Tomb of Queen of Khafre
Pyramid of Menkaure
Mortuary temple Pyramids of queens
Causeway
Valley temple
_ _
470
religion. Toward the end of the Old Kingdom, the construction of smaller pyramids paralleled a decline in royal power. Following the pattern that often occurred in ancient states and empires, the Old Kingdom collapsed after five centuries of strong central rule and was followed by a period of decentralization. The provinces became competitive petty kingdoms that fought with one another, either alone or in small alliances. With the fragmentation of the Egyptian state into these smaller territorial political groups, provincial governors and bureaucrats gained more power. Comparatively large mud-brick tombs were erected in these provinces, displaying the growing power of local authorities. The beginning of the Middle Kingdom, around 2000 B.C., was marked by the reunification of Upper and Lower Egypt under a dynastic line
Chapter 10 States and Empires in Asia and Africa
N
Rock-cut tombs
0
Figure 10.32 The double statue of Menkaure and his sister, Khamerernebti II.
eway
Caus
300 m
centered at the city of Thebes in Upper Egypt. The rulers pacified southern Egypt and then overthrew the dynasty in power to the north. The capital was brought back to Memphis, and the fortified towns of the earlier decentralized period disappeared. With increased administrative centralization, a second era of ostentatious pyramid building began. Yet the Middle Kingdom pharaohs were less despotic. Trading contacts were extended, and parts of Nubia, just to the south of Egypt, were conquered. Huge fortresses were erected to solidify control of this frontier along the Nile. During the first half of the second millennium B.C., central administrative authority again weakened, and a period of short-lived dynasties and regionalization followed. In the sixteenth century B.C., a later Theban dynasty began
pri31057_ch10.pp436-505.qxd
Figure 10.33
10/1/09
8:21 AM
Page 471
Using a shaduf to raise water from the Nile to a walled irrigation ditch.
a third era of unification and centralization, called the New Kingdom. The dynasties of this period are richly documented in historical texts. The kings of the New Kingdom reasserted control from the Nile Delta to Nubia and conquered parts of Lebanon, Palestine, and southern Syria. No earlier Egyptian polity had established such far-ranging external contacts. The highly centralized government of the New Kingdom depended on large external tribute levees for its maintenance and support. During the New Kingdom, kings were considered quasi-divine, the mediators between humans and the gods. These rulers often married their own sisters to concentrate status and divinity in a dynastic line. The kings lived in elaborate complexes that contained throne rooms and extensive private chambers for the royal family, harems, court officials, and servants. The pharaohs adopted new burial customs; their mummies were buried in rock-cut tombs in the Valley of Kings near Thebes. Among the New Kingdom pharaohs were a number of historically renowned figures: Tuthmosis III,
Ahkenaten, Tutankhamen (better known as King Tut), and Ramses II. During the New Kingdom, an important innovation in irrigation permitted the agricultural development of the broad Nile Delta. The laborsaving shaduf enabled one to raise water from wells or ditches onto gardens and fields (Figure 10.33). This innovation, still in use today, may have contributed to the demographic buildup in the delta during the New Kingdom. The kings of the nineteenth dynasty divided their residence between the traditional centers of power, Thebes and Memphis, and a new royal residence that was established in the eastern delta. The administrative centralization and stability that characterized the New Kingdom came to an end around 1000 B.C. Hostilities erupted when strong provincial leaders and local army commanders increased their regional power. A period of foreign intervention followed, and for millennia Egypt did not regain its autonomy for anything but a brief interlude. The Nile Valley was ruled in succession by
shaduf An Egyptian bucket-and-lever lifting device that enables one to raise water a few feet from a well or ditch onto fields and gardens.
_ Site Giza and Dynastic Egypt
471
pri31057_ch10.pp436-505.qxd
10/1/09
8:21 AM
Page 472
Assyrians, Persians, and then Alexander the Great, king of the Macedonians (northern Greece). The almost-3000year succession of pharaohs ended when Alexander appointed one of his generals ruler of Egypt. Several hypotheses have been proposed to explain the highly centralized nature of the early Egyptian state. The constricted distribution and relatively small proportion of usable land (along the Nile) may have fostered the concentration and monopolization of wealth. Commoners and peasants would have had relatively few options when they were heavily taxed or asked to contribute to labor gangs. In an environment such as the fertile Nile Valley, surrounded by desert, it would be very difficult for people to move away and still maintain access to cultivable land.
_ _
472
Chapter 10 States and Empires in Asia and Africa
The critical nature of water control (often requiring centralized maintenance and adjudication) also may have been a factor; however, according to Karl Butzer, of the University of Texas at Austin, early Egyptian irrigation systems were not highly centralized and were largely under local control. Nevertheless, the disastrous potential of Nile flooding may have encouraged the establishment of a central agency to help alleviate periodic local disasters. Finally, Egypt had a number of military-political threats on its borders throughout much of its history, and that, too, may have encouraged the centralization of power in the hands of a few to coordinate a strong defense. Although none of these interpretations is convincing on its own, they do suggest questions and directions for future research.
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 473
Concept
<
Pyramids Ancient monuments across the globe
Pyramid-shaped structures were built by ancient civilizations in both the Eastern Hemisphere and the Americas, representing the wealth and power of those who erected them. Yet the building techniques and the specific uses of these constructions varied greatly from one ancient civilization to another. The best-known pyramids in the world were erected by the ancient Egyptians during the third millennium B.C. The Egyptian pyramids were constructed of cut-stone blocks, with internal passageways and tombs. The first royal pyramid, built around 2680 B . C . at Saqqarah by Zoser, was a sixstep pyramid (see Figure 10.29). Over the next century, the pyramid shape was perfected, culminating in the magnificent pyramids at Giza. The construction of these Old Kingdom pyramids required an enormous expenditure of energy and manpower, involving both unskilled labor to quarry and transport the stones and skilled masons who would cut, fit, and smooth the stones. Given the construction techniques known at the time,
such large structures had to be in the shape of a pyramid to support their weight. Constructed as monuments for kings, the Egyptian pyramids have as their origin the tomb architecture of predynastic Egypt. They were built in one long, contiguous episode of construction, without outside stairways, and were not meant to be climbed once they were completed. The large stone blocks likely were positioned through the use of temporary ramps that were subsequently dismantled. These pyramids served not only as tombs for the pharaohs but also as their houses for eternity. As such, the pyramids reflect the importance Egyptians placed on life in the afterworld. Kurt Mendelssohn argues that the construction of monumental tombs for the pharaohs, who could have been interred at much less cost, was not the only goal. It was the erection of the pyramid itself that was important. He suggests that pyramid construction was a political strategy used by the pharaohs to institutionalize and materialize
AT LAST HAVE MADE WONDERFUL DISCOVERY IN VALLEY; A MAGNIFICENT TOMB WITH SEALS INTACT; RECOVERED SAME FOR YOUR ARRIVAL; CONGRATULATIONS. —Telegram sent by Howard Carter, the investigator who discovered King Tut’s tomb, to his sponsor, Lord Carnarvon, on November 6, 1922
Pyramid of Khufu
Pyramid of the Sun
Figure 10.34 A comparison of the Pyramid of the Sun at Teotihuacan and the Pyramid of Khufu at Giza, which is approximately 146 m high. Concept Pyramids
473
_
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 474
Figure 10.35 The building episodes of a Late Postclassic pyramid in the Basin of Mexico. The earliest pyramid is represented by 0 and the latest by 5. 4 2 5
3
1
3
5
0
y
How did the nature of pyramids and their uses differ between Egypt and Mesoamerica?
_ _
474
the power of the state. The thousands of peasants who worked on the pyramids depended in part on the central government for food during a portion of the year. They were fed from food surpluses obtained from villages through taxation. One consequence was that the population became reliant on the state bureaucracy to redistribute food and organize labor. Recent excavations by Mark Lehner, of Harvard University, and Zahi Hawass have uncovered two communities near the Giza pyramids where builders of these great structures seem to have lived. On the basis of their excavations, Lehner and Hawass conclude that ordinary Egyptians (and not slaves) built the pyramids, some working as conscripts on a rotating basis and others as year-round workers. Monumental structures in the Western Hemisphere were very different in form from their Egyptian counterparts. For example, in western Mexico, such buildings could be circular as well as square and rectangular. Throughout the Americas, most large structures were actually truncated pyramids, with flat tops. Most Mesoamerican pyramids were rubble-filled, with a cut-stone facing. However, some central Mexican pyramids have a core of adobe bricks with a stone facing held together by mortar. Other North American pyramids, such as Monks Mound at Cahokia, were built almost entirely of earthen fill. Mesoamerican pyramids were smaller than the earlier Old Kingdom
Chapter 10 States and Empires in Asia and Africa
pyramids at Giza. Two of the most massive Mesoamerican pyramids, the Pyramid of the Sun and the Pyramid of the Moon, were constructed at Teotihuacan in central Mexico around 100 B.C. At its base, the Pyramid of the Sun is about the size of the Pyramid of Khufu at Giza, but it is only about half as high (Figure 10.34). In contrast to the Egyptian pyramids, Mesoamerican structures were often built in a series of construction episodes, in which they were enlarged and often changed drastically in appearance (Figure 10.35). The largest Mesoamerican pyramid, the Great Pyramid at Cholula (in the state of Puebla), was enlarged many times over hundreds of years. Pyramids in the Americas generally served as foundations for other public buildings, such as temples, shrines, palaces, or elite residences. Stairways provided access to the structures above. The truncated pyramids emulated natural features, such as hills and mountains, raising structures above the ground level. The platforms increased the visibility of the public buildings while making direct access to them more difficult. Although many pyramids in the Americas include burial features, few were designed or used (as were the Egyptian pyramids) exclusively as the final resting place for a particular ruler or elite figure.
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 475
Site
:
An-yang A late Shang city in China
Decades ago, the emergence of Shang civilization in North China appeared to scholars to have no direct antecedents; it was formerly thought that early Chinese cities and states arose as a result of Mesopotamian contact. But as with Egypt and the Indus Valley, more recent archaeological fieldwork has demonstrated that the trajectory from the diversity of the Chinese Neolithic to the rise of early civilizations was more continuous, with outside influences playing a small role. The development of civilization in China was largely indigenous, with its own character and its own writing system. Wheat, barley, and the horse-drawn chariot were significant Western introductions into early China; however, those innovations were not the principal stimuli for the rise of Chinese civilization. Between 5000 and 3000 B.C., the Huang (Yellow) River region of North China was settled by millet and pig farmers who resided in large villages of up to 100 houses. One such village was Ban-po-ts’un (see Chapter 6). The community included large central structures that may have served as clan meeting houses or the residences of certain influential villagers. Marks found on some Yangshao ceramic vessels resemble written Shang characters dating to more than several thousand years later. During the subsequent Longshan period (3000–2205 B.C.), significant changes took place in North Chinese social organization, including marked increases in social ranking. Compared with earlier burials, Longshan mortuary assemblages exhibit more variation; some include jade ornaments and ceremonial weapons. For the first time, many of the Longshan period settlements were walled, and the largest communities were much bigger than ever before.
While the Longshan period was a time of increasing political complexity, population growth, and economic specialization in the Yellow River Basin, similar developments were occurring in other parts of China as well. Thus Chinese civilization appears to have had multiple hearths during the late Neolithic period, with the largest and most complex polities eventually arising in the basin of the Yellow River during the third millennium B.C. During Longshan times, scapulimancy, the interpretation of cracking patterns on heated bone, was practiced (see “Early Writing Systems,” p. 451). Although no written inscriptions are found on Longshan bones, Chinese written characters have been found on numerous bones at later sites (Figure 10.36). In later times, such divination rituals involving oracle bones were addressed to royal ancestors and were carried out by religious specialists. In addition to scapulimancy, the manifestations of ritual were prevalent in many other ways during the Longshan period. Fantastic or mythical animals were crafted on objects of pottery, wood, and jade. Many of these objects were probably associated with shamanistic rituals that were important during the later Three Dynasties and Shang periods. The era following Longshan in North China is known as the San dai, or Three Dynasties: Xia, Shang, and Zhou (pronounced “jo”). The Xia (Sha) dynasty (2205–1766 B.C.) is the first hereditary dynasty in recorded Chinese history. Although later historical texts suggest that there were 17 rulers in 471 years, the period is still somewhat of a mystery archaeologically. K. C. Chang has suggested that the Erlitou culture, with a spatial and temporal distribution that fits the historical accounts of
Figure 10.36 An inscribed oracle bone from the Shang period.
oracle bone An animal bone with
cracks (from heating) or other markings, used to foretell the future.
_ Site An-yang
475
10/1/09
8:22 AM
Page 476
Ri
ve
r
pri31057_ch10.pp436-505.qxd
Hu We
an
i Ri
g
ver
Xianyang
SHANG An-yang Ao/Zhengzhou Erlitou
Lo-yang
Ya
Figure 10.37 The approximate extent of the Shang dynasty in North China, with important sites mentioned in the text.
There is no need to emphasize the significance of Chinese civilization, which produced one of the few pristine states in the world nearly four thousand years ago. But it is rather surprising to note that, compared to other civilizations, little has been done in Chinese archaeology to systematically study the processes of state development. —Li Liu (2004)
_ _
476
ng
e tz
Ri
ve
r
the Xia dynasty, is the archaeological manifestation of that dynasty. The Xia dynasty was transitional between the late Neolithic and the Shang dynasty. Scapulimancy continued to be practiced. Although there are no known inscriptions on oracle bones, signs and symbols have been found on pottery vessels. Bronzeworking, which was in evidence by the end of Longshan, became an increasingly important craft. According to a first-millennium B.C. text, the two principal affairs of the Chinese state were ritual and warfare. Most of the bronze was used to make ritual food and drink vessels, musical instruments, and weapons. Other valuable objects were fashioned from jade, turquoise, and lacquer. A few basic bronze implements were made, but most utilitarian tools, including knives, sickles, and hoes, were made in stone, shell, wood, antler, or bone. Wheelthrown ceramic ware, first manufactured during the late Neolithic, increased in abundance, although handmade pottery also continued to be used. Two palatial house foundations of stamped earth, found in later strata at the site of Erlitou, also indicate a possible important break with Long-
Chapter 10 States and Empires in Asia and Africa
shan. Both palace foundations are much larger than the other houses at the site. The largest foundation, 108 ⫻ 100 m (350 ⫻ 330 ft), was associated with ritual burials (including one individual with bound hands). The decline of the Xia dynasty coincided roughly with the rise of the Shang dynasty (1766–1122 B.C.). The Shang period is known from archaeological excavations at its last capital, Anyang; information from other ancient sites; and written records. The Chinese state had developed by Shang times. The major Shang centers were Ao (an early capital located underneath the modern industrial city of Zhengzhou), Lo-yang, and An-yang (Figure 10.37). Each had a clearly defined ceremonial core, inhabited by a royal household, and a series of nonresidential buildings (meeting halls and ancestral temples). The ceremonial core was surrounded by a service area and the pithouse residences of the commoners. Today Ao is partially hidden by modern buildings, but at its height, it may have covered 3.4 sq km (1.3 sq mi). The central precinct was enclosed by a huge earthen wall 9 m (30 ft) high and 36 m (118 ft) wide at the base (Figure 10.38). The quantity and quality of artifacts indicate the presence of hundreds of skilled craftworkers. One area contained more than a dozen hightemperature pottery kilns, each associated with dense concentrations of oven-fired and broken pottery. In another area, human and animal bones were worked into fishhooks, awls, axes, and hairpins. Bronzeworking was one of the most highly developed crafts. Bronze was fashioned into ornaments as well as tools. In one workshop, molds were used to mass-produce bronze arrow points. As yet, the Shang materials contemporaneous with Ao have yielded no definitive indications of writing, chariots, or royal mausoleums, all features solidly linked with Anyang and its late Shang materials. However, the construction technology and the basic layout of the palace structure at Zhengzhou were nearly identical to those at subsequent An-yang (Figure 10.39).
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Toward the end of the Shang dynasty, the capital was moved north to An-yang (ahn-yong). Like the earlier capital, An-yang was a large ceremonial and administrative center with monumental architecture surrounded by craft areas, including bronze foundries, stone and bone workshops, and pottery kilns. Circling the center were residential hamlets (small concentrations of wooden houses with thatched roofs), a royal cemetery area at Hsi-pei-kang, and more workshops. An-yang consisted of three groups of buildings with a total of 53 rectangular structures built on top of stampedearth platforms. The largest structure was 60 m (195 ft) long and is presumed to have been a royal palace. Between two of the building groups is a square, earthen foundation thought to have been a ceremonial altar. As a whole, this building complex appears to have been well planned, although it almost certainly was implemented in several construction stages. Late Shang society was highly stratified into upper and lower classes. The extent of social distinctions, present as early as the Longshan period, was exaggerated by the establishment of An-yang. The king, his family, and officials were at the top. Kings were considered divine, with power flowing from the king to the nobility to the court, and finally to the commoners. Only those of high status possessed the spectacular Shang bronzes, used the Chinese script, and controlled the archives. The king and his court received grain and other forms of tribute, which they used to support a lavish style in death as well as in life. In the 11 large royal tombs at Hsi-peikang, Shang kings were buried with sacrificed retainers, horse-drawn chariots, and large quantities of luxury items, including bronze vessels, shell and bone ornaments, jade, and pottery (Figure 10.40). For each of these graves, moving the earth alone would have required thousands of working days. The royal tombs were surrounded by 1200 smaller and simpler graves, most of which lack any grave goods.
Page 477
Bone workshop Bronze foundry
Chin-shui Rive
r
Pottery kilns
Cemetery area Cemetery area
N
City wall aboveground City wall underground
Hsiung
-erh Riv
Wall gaps Stamped earth foundations 0
Cemetery area
500 m
er
Cemetery area
Bronze foundry
Figure 10.38 The central precinct of Ao, enclosed by a huge earthen wall. Cemeteries and craft workshops have been found in many areas outside the city wall.
The lower class consisted of farmers and craftworkers. Some of the more skillful and highly specialized ones (bronzesmiths, lacquerwarers, and wood carvers) may have had some of the privileges of the upper class; yet overall, the farmers and craftspeople used stone and bone implements and coarse gray pottery and lived in semisubterranean pithouses. The commoners labored for public works and
Figure 10.39 An artist’s reconstruction of a Shang palace at Ao.
_ Site An-yang
477
pri31057_ch10.pp436-505.qxd
Figure 10.40 at An-yang.
_ _
478
10/1/09
8:22 AM
Page 478
Sacrificial burials
military campaigns. At the bottom of the lower class were the war captives, who were kept as slaves or served as sacrificial victims for rituals and temple dedications. Shang civilization is famous for its bronzework (Figure 10.41). Food and drinking vessels are the most common bronze items, but some weapons, chariot and cavalry fittings, and musical instruments also were made of bronze. Small objects, such as spearpoints, were made by pouring a molten mix of copper and tin into molds. Large ceremonial vessels were made by a more complex process involving clay prototypes. During the first half of the twentieth century, Western scholars familiar with the An-yang bronzes presumed that the metalworking technology was introduced from Europe or Southwest Asia. Yet the absence of such Western techniques as annealing, hammering, and lost wax casting and the presence of complex mold-casting technologies indicate an indigenous origin for Shang bronzeChapter 10 States and Empires in Asia and Africa
working. The roots of late Shang metallurgy may lie in the sophisticated pottery kilns that were used as early as the late Neolithic. At An-yang, scapulimancy became more complex and sophisticated. In addition to the shoulder blades of cattle and water buffalo, diviners used the carapaces of turtles. Once an answer was obtained, by consulting ancestral spirits and interpreting the cracks on the oracle bones, both the question and the answer were sometimes recorded on the surface of the bone. A symbol similar to the later written character for “book” is found on one bone, indicating the presence of scribes. The earliest Chinese writing may have been expressed on silk, bamboo, or wooden tablets. These surfaces commonly were used for written inscriptions in the last half of the first millennium B.C.; however, no texts on these highly perishable materials have been found for Shang. By late Shang times, Chinese written language had developed to the point where over 3000 phonetic, ideo-
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
graphic, and pictographic symbols were in use. Based on more than 150,000 inscribed turtle shells, in addition to inscriptions made in bronze, pottery, and stone, it is clear that early Chinese writing was related closely to the political, military, and ritual activities of the upper class and had little to do with mercantile matters. The borders of the Shang state are unknown; however, late Shang rulers had at least some control over a fairly large area in northern China. The extent of their influence varied according to distance from the capital. Shang rulers traveled widely across their domain, and the extent of their influence was related in part to their actual physical presence. The rulers were assisted by a complex hierarchy of local nobles, who had considerable autonomy in their own territories. These local lords were responsible for collecting taxes and supplying men for public projects and military campaigns. On occasion, armies numbering 30,000 soldiers were assembled to wage war against “barbarians” at the edge of the Shang domain. Military success depended on the horse-drawn chariot. War was waged more for people than for land. In these campaigns, thousands of prisoners were taken, most of whom were sacrificed or used as slaves. Despite major changes in political organization, written communication, and social stratification, the basic subsistence technology of the late Shang period changed little from earlier times. Millet remained the principal food crop in North China, supplemented by rice and wheat. Stone hoes, harvesting knives, and wooden digging sticks remained the primary cultivation implements. In some areas, two crops a year were grown, suggesting irrigation, yet large-scale water control was not implemented until the last centuries B.C. Changes in labor practices may have constituted the most dramatic shift in the Shang economy. Larger numbers of people were employed in farming, thereby increasing production per unit of land as well as the amount of cultivated terrain. The reliance on human
Page 479
energy may help account for the kind of military campaigns that were waged, as well as the rapid demographic increases that occurred in China during this period. The importance of agricultural labor may have encouraged rural families to grow, since children could work in the fields at an early age. One consequence would have been large-scale population growth. The Shang dynasty was overthrown by people living on its western periphery in the vicinity of Xianyang on the Wei River. It is unclear whether the defeat resulted from the internal rebellion of a distant Shang province or by the rebellion of an outside group of people forced by nomadic pressures to settle in or near the Shang state (as suggested by one 3000-year-old text). Whatever the case, the new dynasty did not create an entirely new civilization; it incorporated the existing network of towns and officials. It was on the foundation of the Shang dynasty that the subsequent Zhou dynasty established China’s first empire.
Figure 10.41 A bronze vessel from the Shang or early Zhou period.
Western scholars tend to see civilization as a level of human attainment that lifts humans to a higher plane than that of “mere” animals and plants, and as an artificial environment that insulates us from unadulterated and hostile nature. But in ancient Chinese civilizations humans and nature were regarded as one; the Chinese were civilized precisely because they were able, or at least desirous, to be close to and harmonize with nature as a matter of conscious and deliberate choice. —Kuang-Chih Chang (1994)
_ Site An-yang
479
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 480
Concept
<
The Roots of Chinese Cuisine Ancient culinary ritual and traditions
Figure 10.42 Late Neolithic noodles from China, revealed after the inverted earthenware bowl that had contained them was removed. These spectacular remains were found in northwestern China at the upper reaches of the Yellow River.
_ _
480
Archaeologists often study containers of fired clay, stone, and metal to infer information about chronology or past technologies. However, the main function of these vessels was the preparation and serving of food and drink. The pottery and bronze vessels of the Shang and Zhou dynasties provide one perspective on ancient Chinese culinary traditions. This vantage recently has been supplemented by the recovery of actual food remains, such as the incredible discovery of millet noodles that were found in a sealed earthenware bowl and date back roughly 4000 years (Figure 10.42). Historical texts and inscriptions also yield key insights on ancient Chinese cuisine. For example, The Three Lis, or Three Books of Rites, solemn texts from the Han period (206 B.C.–A.D. 220) that record ritual and courtly behavior from ancient China, provide detailed descriptions of food preparation and feasts, including the types and amounts of food and wine served on certain occasions. Oracle bone and bronze vessel inscriptions from the Shang period include written characters that refer to certain foods, rituals, and cooking techniques and whose shapes suggest ancient practices. It is not surprising that many groups of people are strongly concerned with food. However, according to K. C. Chang, the Chinese are more concerned with food and eating than any other people. He pointed out that in contemporary China, a familiar greeting is “Have you eaten?” Chang believed that this preoccupation has a long history, as do many Chinese culinary traditions. According to one account, when a duke asked Confucius (551–479 B.C.) about military strategy, the sage replied, “I have indeed heard about matters pertaining to meat stands and meat platters, but I have not learned military matters”
Chapter 10 States and Empires in Asia and Africa
(quoted in Chang, 1973, p. 496). In ancient China, knowledge and skill in the preparation of food and drink were important personal characteristics; King T’ang, the founder of the Shang dynasty, chose a cook as his prime minister. According to several Zhou texts, a cooking vessel, called the ting cauldron, was the primary symbol of the state. Along with differences in dress and adornment, part of the definition of a “barbarian,” or a non-Chinese, was a person who did not prepare and consume food in the customary manner. Included in the Three Books of Rites is a personnel roster of the king’s palace; almost 60% of the 4000 people responsible for the king’s private residential quarters were involved with food and wine, with specialists assigned to menu preparation, service, meat, fish, game, shellfish, wine, fruits and vegetables, pickling, and salt. The dualism between yin and yang, central to much of contemporary Chinese culture, is also present in the categorization of food. A basic dichotomy exists between food and drink. Food is further divided into grain food and “dishes” that are combinations of meat and vegetables. These distinctions, and their associated rules, are integral to the Chinese way of eating today and seem not to have changed much since Zhou times. In the past as well as today, the basic or essential meal consists of grain and water. The basic word shih, for food (as opposed to drink), also, in a narrower sense, refers to grain. Meat and vegetable dishes are considered secondary in importance and are to be eaten in moderation. The historical texts indicate that eating together was a major source of enjoyment in ancient China. Yet it also was a serious social affair, with strict rules to be observed, including correct table manners and an etiquette regarding the appropriate foods to be served
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
at specific occasions. In Shang and Zhou times, people from the upper ranks ate individually, kneeling on mats. The number of dishes served to an individual was primarily determined by one’s rank and age. Utensils and serving platters were placed beside the individual in specific arrangements, with certain foods on the right and others on the left. In addition, rules specified the arrangement of certain foods on the serving vessels and the way they were presented at the table. Children were trained early to eat with their right hand. In late Zhou times, crunching bones with one’s teeth, eating too quickly, slurping down soup, and picking one’s teeth were all specified as inappropriate etiquette. In one contemporary Chinese cookbook, 20 methods for heating food are mentioned. Many of these procedures, including boiling, steaming, and roasting, also were important in Zhou texts. Yet the ancient accounts do not mention stir-frying, which is prevalent today. Nevertheless, in the past as well as the present, it is the preparation before cooking that is essential to Chinese cuisine. The word for cooking in the Zhou texts literally means “to cut and cook.” Great significance is placed on the art of mixing flavors and ingredients into distinctive soups and stews before they are heated. Social class was a major factor in dietary variation. Peasants relied on a basic grain diet, using the platters and vessels associated with such foods. According to the textual record, the meat dishes were intended for rituals or upper-class feasts. Meat was usually dried, cooked, or pickled. A few recipes for elaborate dishes of the Shang and Zhou times are known; these are for the so-called Eight Delicacies, which were prepared specifically for the elderly. They included the Rich Fry (pickled meat over rice), the Similar Fry (pickled meat over millet), the Bake (baked stuffed pig or ram), the Pounded (pounded meat fillets softened by pickle and vinegar), the Steeped (newly killed beef steeped in wine), the Grill (dried meat seasoned with cinnamon, ginger, and salt), the
Page 481
Soup Balls (fried cakes of meat and rice), and the Liver and Fat (roasted dog liver cooked in fat). In ancient Chinese cooking, each variety of food and drink was associated with different ceramic and bronze vessels (Figure 10.43). There were vessels for boiling, for simmering, and for steaming. For serving and eating, ladles and chopsticks were used, although hands were used as often as chopsticks. There were special cups for water and wine. In the era of the Shang dynasty, bronze vessels were used to serve grains and drinks made from grain, but never to serve meat. The most important vessels for meat dishes were made from wood, basketry, and pottery. Most excavated Zhou period burial assemblages contain ceramic vessels. The majority of the graves include not just one or two containers but a whole range of vessel forms. Chang reasoned that most of these individuals were interred with a set of containers for cooking and serving grain, serving meat, and drinking. Chang’s deduction would have been difficult, if not impossible, to derive without the careful juxtaposition of texts with archaeology. Together, these complementary records are serving to uncover the deep traditions of Chinese cuisine.
Figure 10.43 Examples of Shang and Zhou food and drinking vessels: drinking vessels (top), serving vessels (center), cooking vessels (bottom).
Chinese food is one of the great cuisines of the world. Discuss its roots and its impact on material culture.
Concept The Roots of Chinese Cuisine
y _
481
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 482
Site
:
Xianyang Terracotta soldiers and the Qin dynasty
Figure 10.44 A reconstructed segment of the Great Wall of China in the mountains north of Beijing.
The Zhou dynasty (1122 to third century B.C.) marks the beginning of imperial China and its traditions, which persisted for the next 2000 years and into the present. Zhou society was highly stratified at its center, with the king and a royal court at the top. Away from this core, the adjacent areas were divided into partially independent provinces, and administration was enacted by semifeudal lords who had great control over their local domains. Periodic civil wars erupted between these lords and the king. The Chinese state during the Three Dynasties (including the early Zhou)
_ _
482
Chapter 10 States and Empires in Asia and Africa
was built on a hierarchical network of large lineages in which the distance away from the main male line of descent determined relative political status and access to power. Each walled town was inhabited primarily by members of a particular lineage. The latter half of the Zhou period was characterized by great political change and upheaval, with warring states and shifting capitals. It also was a time when Chinese urbanism spread over a much wider area than ever before. Great cities were built, many of which were larger and more nucleated than the earlier Shang cities. The largest
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 483
Figure 10.45 An artist’s rendering of the gallery of terracotta soldiers and horses guarding the east gate of the emperor’s tomb at Mount Li.
Zhou settlement, G’a-to, had 270,000 people. All the large cities were walled. By 600 B.C., iron casting was practiced and iron agricultural tools were in use. Large irrigation works were constructed, and wet-rice irrigation became increasingly important. Changes in agricultural technology enabled rapid increases in population density. Late Zhou socioeconomic structure placed great emphasis on the taxation of peasants in lieu of labor drafts. Kinship bonds began to diminish, and territorial units and bureaucracies gained importance. Late Zhou was the time of Confucius (or Kongzi in Chinese), who preached order, deference, and family ties, perhaps in response to rapid social transition and transformation. Although large-scale political integration remained relatively weak and fragmentary, a single system of measurement was adopted across most of China. There was increased interregional trade and commercial activity, as well as greater cultural unity.
By the third century B.C., the descendants of the western Zhou kings ruled an increasingly small area outside their original homeland. As the Zhou polity weakened, other states rose in influence. The Qin polity expanded, and its short-lived dynasty (221–206 B.C.) eclipsed the Zhou, along with five other contemporary states. Ying Zheng inherited the throne of the Qin (pronounced “chin”) kingdom at age 13 in 246 B.C. During the first 25 years of his reign, he frequently engaged in battle, eventually conquering six other major kingdoms. For that reason, six was considered the lucky number of the Qin. Through military prowess, he unified China into a single imperial kingdom in 221 B.C. and declared himself China’s first emperor, taking on the name Shihuangdi (literally, “first august emperor”). The empire was ruled from the capital city of Xianyang (she-ON-yong), to which he forced over 100,000 royal and wealthy families from throughout the empire, to
_ Site Xianyang
483
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Figure 10.46 Rows of terracotta soldiers in the large rectangular gallery guarding the emperor’s tomb.
The First Emperor of China controlled a vast territory and wielded enormous power. He ordered 120,000 families to move to the new capital, Xianyang; he summoned 700,000 men to build his tomb and other structures; and he was self-consciously aware of his authority and of the new era that this marked. Long inscriptions carved at his command on mountains in eastern China described his achievements and proclaimed his universal, indeed cosmic, rulership. —Jessica Rawson (2007)
_ _
484
Page 484
move. Shihuangdi had luxurious palaces built in Xianyang that were replicas of royal residences in the conquered states. By moving local lords to Xianyang, he forcibly detached the feudal aristocracy from the land and its people, weakening their power. This move also served to centralize the Qin empire by concentrating economic and political power in a single capital. According to historical records, Shihuangdi was an ambitious and ruthless emperor. He built the Great Wall along China’s northern periphery by joining walls that had been constructed by earlier feudal states (Figure 10.44). Although the traditional view is that the wall was intended to protect the newly formed empire from the nomadic herders of Asia to the north, other scholars have suggested that its main function was to prevent heavily taxed peasants from escaping
Chapter 10 States and Empires in Asia and Africa
taxes and conscription. The 2400-km (1500-mi) wall, built by 700,000 conscripts and wide enough for six horses abreast, remains the longest fortification anywhere. Many men perished while working on the wall, inspiring some to call it “the longest cemetery in the world.” Shihuangdi also established China’s first standing army, a body that may have contained more than a million people. To weaken regional autonomy, Shihuangdi destroyed the feudal structure that had existed for centuries. Because he saw Confucian philosophy as a threat to his authority, all the books of this school were burned, and Confucian scholars who refused to accept his reforms were buried alive. The centralizing tendencies of Shihuangdi included increasing codification of a Chinese legal system and the standardization of Chinese character
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 485
Figure 10.47 Workers measuring a terracotta soldier guarding the tomb of Shihuangdi.
writing so that the written language could be understood throughout the empire. Weights and measures, coins, and the gauges of chariot wheels were increasingly regulated and made more homogeneous. Paper was invented during the Qin dynasty. In the grave of one Qin official, more than 1200 bamboo slips were found, bound into a series of books and containing an explicit legal code specifying particular crimes and their punishments. Under Shihuangdi, road building was intensified and a canal system was constructed to enhance communication and transportation. The canal system was one of the greatest inland water communication systems in the ancient world, and several canals are still functioning today. As soon as Shihuangdi became emperor, he began building his tomb. According to history, 700,000 laborers
from all parts of the country worked for 36 years on the project, a virtual subterranean palace for the emperor to live in for eternity. Recent DNA analysis has documented diverse origins for the workers. According to early Chinese records, the architects of the tomb conceived of it as a universe in miniature. All the country’s major waterways were reproduced in mercury within the tomb, and they fed into a tiny ocean. Heavenly constellations were painted on the ceiling. The emperor’s outer coffin was made of molten copper, and fine vessels, precious stones, and other rarities were buried with him. The burial tomb, called Mount Li, was at one time 46 m (150 ft) tall. Built in the center of a spirit city, an area enclosed by an inner wall, it contained sacred stone tablets and prayer temples. Beyond this area was an outer
_ Site Xianyang
485
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 486
Figure 10.48 A life-size terracotta figure from the gallery.
city enclosed by a high rectangular stone wall 7 m (23 ft) thick at the base. The total complex covered 200 ha (500 acres). Today, most of the walls and temples have been removed. About 1370 m (4500 ft) east of Mount Li, excavations have revealed one of the most astonishing ancient spectacles. Guarding the east side of the emperor’s tomb is a brick-floored, 1.2-ha (3-acre) gallery of terracotta soldiers and horses (Figure 10.45). Collapsed pillars indicate that a roof once covered the underground battlefield. In the royal tombs of the previous Shang dynasty, kings and high-ranking officials were interred with living warriors, women, servants, and horses. This practice, which had ceased cen-
_ _
486
Chapter 10 States and Empires in Asia and Africa
turies before the Qin dynasty, evidently was revived in symbolic fashion by Shihuangdi. Although only parts of this large rectangular gallery, and two nearby smaller ones, have been excavated, some 8000 terracotta figures have been exposed, along with wooden chariots (Figure 10.46). The terracotta warriors are slightly larger than life-size (Figure 10.47); they are arranged in battle formations, dressed in uniforms of various rank, and carry real weapons—swords, spears, and crossbows. Traces of pigment indicate that the uniforms were brightly colored. Of the excavated figures, no two look exactly alike; their facial expressions vary, suggesting that they were realistic portraits of each in-
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 487
dividual in the emperor’s honor guard (Figure 10.48). Even the horses were very finely crafted, appearing alert and tense as they would be in battle. The names of more than 80 master craftsmen, drawn from imperial workshops as well as other parts of China, have been identified on the backs of figures in the large gallery. The army and horses are supplemented by a rich artifact assemblage, including gold, jade, and bronze objects, linen, silk, bamboo and bone artifacts, pottery utensils, and iron agricultural tools. Elemental analysis of the swords has revealed that they were made from an alloy of copper, tin, and 13 other elements. The designers of the tomb’s security system, a series of mechanized crossbows, were sealed inside the tomb to die so that none of the tomb’s secrets could be divulged. Shihuangdi always lived and worked in guarded secrecy, because several assassination attempts were made on his life. Only a few trusted ministers ever knew where he was. He died on a journey to the eastern provinces, and his death was kept a secret from all except his youngest son. His prime minister and his chief eunuch (a castrated male) apparently plotted to keep the death secret for their own ambitious reasons. They wanted the emperor’s youngest son to succeed to the throne, instead of an elder son, as Shihuangdi had decreed. The councillors thought they could more easily influence and manipulate the younger son. The elder son, ex-
iled to the northwestern frontier to help build the Great Wall, was sent a fake order to commit suicide, which he did, paving the way for the younger son to become the new emperor. Nevertheless, Shihuangdi’s efforts to expand his domain to both the north and the south sapped his treasury, so Qin preeminence was short-lived. Although the Qin dynasty was brief, China’s first episode of unification was not. Qin rule was followed by the Han dynasty, which lasted for 400 years (206 B.C.–A.D. 220). The Han unification was made possible in part by technological innovations developed in Zhou times: iron tools, wet-rice irrigation, the ox-drawn plow, improved roads, and the crossbow. Under Han rule, China continued as a unified empire, but with greater political stability. The economy was prosperous, and a standardized coinage circulated throughout China. During the Han dynasty, China became even more densely settled. The world’s first census, in the years A.D. 1 and 2, lists the population of the empire as 57.7 million, with cities of up to 250,000 people. One late Han city may have contained as many as 500,000 people. The decisions made by the Han monarchs were implemented through 1500 administrative provinces, each of which was centered at a walled town. No other political system of its era—not even the Roman Empire—was as vast in size or bureaucratic complexity.
_ Site Xianyang
487
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 488
Site
:
Angkor Maritime kingdoms in Southeast Asia
MYANMAR LAOS
THAILAND
M
. iR Ch R. un M
R. ng e ko
Bang Pakong R.
Khok Phanom Di
VIETNAM
Angkor CAMBODIA
Gulf of Thailand 0
Figure 10.49 Southeast Asia, with the location of Angkor, the Khmer ceremonial center situated in the jungles of Cambodia along a tributary of the Mekong River.
_ _
488
300 km
Until the introduction of cultivated rice in the third millennium B.C., Southeast Asia was populated by huntergatherers who lived in small communities along the coast and in forested uplands. At coastal sites such as Khok Phanom Di (see Chapter 6), villagers took advantage of rich coastal resources and rivers that facilitated exchange with distant regions to establish year-round settlements. With the introduction of cultivated rice from South China, small agricultural villages were settled throughout the area. Communities such as Khok Phanom Di engaged in exchange with the inland farming communities and soon cultivated rice themselves. Rice rapidly became the staple food crop across Southeast Asia. Between 1500 and 1000 B.C., bronze casting was adopted in Southeast Asia, and after 500 B.C., iron was smelted to
Chapter 10 States and Empires in Asia and Africa
make ornaments, weaponry, and agricultural tools. Settlements began to grow in size, and more hierarchical forms of leadership and socioeconomic stratification were evidenced. Before the end of the first millennium B.C., Southeast Asians began to engage in maritime trade that linked them to offshore islands and China, as well as to the Indian subcontinent. With the introduction of new technologies and belief systems, some Southeast Asian societies became highly ranked kingdoms, focused on large centers that were presided over by an aristocratic class with close relationships to the ancestors. Ruling power was based in part on control of agricultural land, rice surpluses, and advantages in access to high-status imported goods. Public displays of feasting and ritual were important aspects of maintaining power and legitimacy to rule. For hundreds of years, the landscape of Southeast Asia was peopled by small, competing polities that were often in a state of political flux and whose boundaries frequently shifted over time. Warfare is a recurrent theme in inscriptions, which describe military expeditions and the conquest of rivals. Social relations revolved around rulers, whose personal and spiritual qualities determined their success in forming alliances and defeating rivals. Although belief systems were borrowed from South Asia, the distinctive form of kingship in Southeast Asia was markedly different from that of the much earlier Indus civilization, which did not emphasize specific individuals through elaborate graves or personalized inscriptions. Rather, one can draw parallels between Southeast Asian kings and the Classic Maya noble lords of Mesoamerica. As with the Classic Maya, Southeast Asian kingship emphasized their divine monar-
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 489
Figure 10.50 The large ceremonial complex at Angkor. Rectangular reservoirs and ancient canals are visible in this aerial view. The large square to the right of the large reservoir is the city of Angkor Thom, with the smaller complex of Angkor Wat just below it.
chy and fostered the concentration of wealth and luxury in their hands. After A.D. 550 in the Mekong Valley region of Cambodia, a series of kings tried to establish hegemony over as large an area as possible, but for several hundred years none was able to hold the kingdom together for long. The Angkor (AHN-cor) state that arose on a tributary of the Mekong River just after A.D. 800 was one of the largest and most centralized of these Southeast Asian polities (Figure 10.49). The formation of this large, centralized state in A.D. 802 is attributed to Jayavarman II, a dynamic Khmer ruler who joined a series of smaller competing polities into a large state by first defeating rival rulers and then placing his followers in positions of authority. He established a succession of Khmer dy-
nasties whose reign at Angkor endured for more than 600 years, in spite of periodic disruptions caused by civil wars and conflicts over succession. Angkor, derived from the Sanskrit word for “holy city,” is the modern name for a large complex of monuments, temples, reservoirs, and walls that was the political and ceremonial center of the Khmer civilization (Figure 10.50). The more than 100 temples at the site were constructed by a series of Khmer dynasties between the ninth and fifteenth centuries A.D. At its height, in the twelfth century, the site of Angkor stretched more than 1000 sq km (600 sq mi) and may have had between 500,000 and 1 million residents. Majestic temples were an important symbol of rulership, and each new king built a massive religious structure
_ Site Angkor
489
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Figure 10.51 Angkor Wat, the largest religious monument in the world. It was constructed as a representation of the Hindu universe by King Suryavarman II, who ruled from A.D. 1113 to 1150. Almost every surface of the temple and walls is covered with elaborately carved scenes.
_ _
490
Page 490
to commemorate his reign. At his death, the temple became the deified ruler’s mausoleum. The largest and most elaborate temple at Angkor, Angkor Wat, was constructed by Suryavarman II soon after he became king in A.D. 1113 (Figure 10.51). The construction of the temple may have taken the greater part of his 37-year reign. The walled complex measures 1500 m (5000 ft) by 1200 m (4000 ft) and incorporates a large moat, 200 m (660 ft) wide, that surrounds a second walled enclosure. The central precinct inside the smaller enclosure measures 215 ⫻ 187 m (710 ⫻ 620 ft) and includes three raised, covered terraces that culminate in a monument with five towers, the tallest of which stands 65 m (215 ft) above the surrounding terrain. These sandstone towers were carved in the form of giant lotus buds. The walls of the complex are covered with elaborate scenes in bas-relief that depict the king and his court, processions, and battle scenes. Scenes on later temples portray activities of everyday life, including cooks preparing food for a feast, a wealthy Chinese merchant entertaining guests, men fishing, women selling fish in market stalls, and men playing chess. Many Angkorian inscriptions detail the punishments meted out to nonbelievers and the heavenly
Chapter 10 States and Empires in Asia and Africa
rewards awaiting the faithful. Small inscriptions on walls describe the punishment for particular crimes, including greed or gluttony; the theft of land, animals, and other goods; and inciting opposition. Temple inscriptions also reveal an intense concern with the quality of agricultural land and details of ownership and boundaries. Construction of the temples and other public works, including the large reservoirs at Angkor, required enormous labor forces. Local labor was tied to the temples and provided them with goods and labor for rituals and for maintaining the ruling elite. The Khmer civilization had no currency; through taxation, the royal court received and distributed huge amounts of subsistence goods including rice, fish, salt, honey, oil, and cloth. The Khmer rulers headed a bureaucracy of high-status families that included royal councillors, generals, and administrators. The bureaucracy controlled most aspects of Khmer life. There were numerous grades of officials, who oversaw royal warehouses and the payment of taxes, monitored land boundaries, commanded the army, and organized labor for construction projects. The economy of the Angkor and other Southeast Asian states was based
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
on surplus rice agriculture, and some scholars have argued that the huge reservoirs at Angkor were constructed for irrigating rice fields. The reservoirs are very large, holding millions of cubic meters of water, and are connected to canals and moats. But others disagree, arguing that there is plentiful groundwater in the region, so that irrigation was not necessary to produce rice surpluses. Inscriptions that mention the reservoirs do not link them with irrigation, and the reservoirs instead may have been constructed to provide drainage and help control floods. They also may have had symbolic importance, representing lakes surrounding sacred sites in the cosmos. Their very size may have been a measure of the power of the kings who constructed them. During the twelfth century, Angkor was sacked by a rival. The Khmer rebuilt
Page 491
the center under the last of their major rulers, Jayavarman VII, who constructed the largest ceremonial precinct at the site, Angkor Thom, during his reign. This large precinct was surrounded by a massive wall 12.8 km (8 mi) long, entered through large, ornately carved gates 23 m (75 ft) tall (Figure 10.52). Eventually, warfare became endemic, and Angkor was abandoned in A.D. 1431 after it was sacked by the Thai following a long siege. After its abandonment, Angkor was visited by many explorers. By the end of the nineteenth century, clearing the jungle from the temples had begun, and restoration and maintenance continues to the present. Although the vegetation, war, and looting have taken their toll, even today the grandeur of Angkor inspires visitors, both Khmer and foreign, and archaeologists from around the globe.
Figure 10.52 The ornately carved southern entrance gate to the city of Angkor Thom, topped with five stone heads.
_ Site Angkor
491
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 492
Site
:
Jenné-jeno An ancient urban center in West Africa The ancient Egyptian state, with its close ties to the Mediterranean world, had few direct contacts with regions far to the south and west. Only the area down to the Nile’s first rapids (near the present Aswan Dam) was ruled consistently by the pharaohs, although slaves, ebony, and ivory were brought from farther south through trade. Early in the last millennium B.C., following New Kingdom penetrations into Upper Nubia, the Nubian king-
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on Jenné-jeno, see the Internet exercises on your online learning center.
Figure 10.53 Africa, with places and names mentioned in the text.
Medi terran ean S ea
Ni le
S a h a r a
Aswan
D e s e r t
er Riv
Upper Nubia
Timbuktu er N ig
Jenné-jeno
Meroë
Ri ve r
Indian Ocean
Atlantic Za m b
Ocean
ez i
Karanga empire Limpopo River
0 0
500 500
River
Zimbabwe Mapungubwe
1000 Miles
1000 Kilometers
_ _
492
Chapter 10 States and Empires in Asia and Africa
dom of Kush (in what is now the Sudan) arose (Figure 10.53). The Kushites were ethnically and linguistically different from the Egyptians, and they had their own tomb style. Egypt later came briefly under Kushite control, and the intermingling of these two East African cultures intensified. During the fourth century B.C., following an episode of Egyptian control, a major center on this portion of the southern Nile was established at Meroë. For almost a millennium (until the fourth century A.D.), the Meroitic kingdom maintained trade connections with both the Mediterranean world and the southern Sahara. Yet, after its collapse, the fertile grasslands around Meroë became the homeland for more rural lifeways. To the west, state-level societies did not develop on the southern fringes of the Sahara until the first millennium A.D. The Sahara Desert has undergone several cycles of hyperaridity (extreme drought) that have affected human settlement in the region. Before 10,000 B.C., the desert was very dry and uninhabited. Several millennia later, after conditions improved, the Sahara consisted of a mosaic of shallow lakes and marshes linked by permanent streams where communities of hunter-fishercollectors settled. The technology of these early groups included microliths, bone harpoons, comb-impressed pottery, and grinding stones used to process wild grains. Livestock were added to the subsistence base in the fifth millennium B.C. Sheep and goats were introduced first, most likely from Southwest Asia via Egypt or from North Africa. It is unclear, however, whether domestic cattle also were introduced from Southwest Asia or were domesticated indigenously from a wild local species.
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
The central and southern Sahara continued to be occupied by pastoralists until about 2500 B.C., when a second pronounced dry period began, which continues to the present. Desiccation of the environment led to the southward movement of the Saharan peoples into sub-Saharan Africa. These herder-collectors brought their cattle with them, while the people remaining in the Sahara became more nomadic. The presence of the tsetse fly in the Sahel (the region at the southern edge of the Sahara) and the Sudanic savanna before 2500 B.C. had limited the expansion of pastoralists into the region. With the onset of drier conditions, the tsetse fly migrated farther south. By 1000 B.C., pastoralists of the western Sahel had adopted or domesticated cereals. Early West African staples included sorghum and several genera of millet. Yet the subsistence economy remained variable as wild foods continued to be important, and lifeways based on hunting, gathering, and fishing remained dominant in some areas. Iron metallurgy was introduced into West Africa during the first millennium B . C . In much of Africa (except Egypt), iron was the first metal to be used. This pattern contrasts with the development of metalworking in most of Eurasia, where ironworking followed copper and bronze. One explanation is that ironworking was introduced from the north, most likely by iron-using Phoenicians who established trading colonies in North Africa. Yet there is limited evidence for copper smelting in the West African Sahara almost a millennium earlier that may have led to the independent discovery of iron metallurgy in West Africa. Whatever its source, the great efficiency of iron tools spurred the rapid spread of this technology throughout Africa on its introduction. The earliest identifiable iron-using society in West Africa is the Nok culture of central Nigeria, known for its distinctive sculpted terracotta heads. Iron-smelting furnaces from the region
Page 493
Figure 10.54 Jenné-jeno, located in the floodplain of the Niger River.
date to the last half of the first millennium B.C. One of the few other areas in West Africa that had iron at that time was southwestern Mali, where iron slag and artifacts have been recovered from the earliest occupation levels at the classic tell site of Jenné-jeno. Situated 3 km (1.8 mi) southeast of the present city of Jenné, the site consists of a mound of successive settlements 800 m (2600 ft) long and 6–8 m (20–26 ft) high (Figure 10.54). Jenné-jeno is located in the Inland Niger Delta in southwestern Mali along the middle course of the Niger River, an area lined with hundreds of ancient tells. This vast inland area of swamps and standing water is remarkably fertile, and the rivers have a great abundance of fish. The alluvial plain, however, is devoid of stone, copper, and iron ore. The initial settlement at Jenné, dating to around 200 B.C., was confined to a few hectares in the central part of the site. The early inhabitants were mixed agriculturalists who constructed circular houses of straw coated with mud. The earliest direct evidence for domesticated African rice (Oryza glaberrima) in West Africa, dating to the first century A.D., was recovered at Jenné-jeno by the site’s excavators, Susan and
pastoralist An animal herder; pastoralism is a subsistence strategy generally associated with a mobile lifeway.
_ Site Jenné-jeno
493
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Figure 10.55 A funerary urn uncovered in the cemetery at Jenné-jeno.
_ _
494
Page 494
Roderick McIntosh, of Rice University. Fish, reedbuck (African antelope), and domestic cattle (Bos taurus) also figured heavily in the inhabitants’ diet. Craft skills are evident in the earliest materials recovered at the site, which include iron and slag and significant quantities of well-manufactured ceramics decorated with twine impressions. Jenné-jeno’s strategic location along a navigable river in a floodplain lacking raw materials encouraged exchange. The villagers traded their agricultural, fish, and animal products for iron ore and grindstones from at least 50 km (30 mi) away. The development of these regional trade networks within the Inland Niger Delta and immediately adjacent areas was an important factor in Jenné’s early expansion. The practice of camel domestication, which first occurred in Arabia, spread to the Sahara by early in the first millennium A.D. Camel transport
Chapter 10 States and Empires in Asia and Africa
had an important effect on the economy by providing a means of regular long-distance trading across the desert. With the development of long-distance trade, villages such as Jenné-jeno became important market centers. Jennéjeno expanded its trading network to include copper and salt from the Sahara to the north and gold from the savanna and forest country to the south. Jenné-jeno became increasingly urban after the mid-fourth century. The most frequent mode of interment in the crowded cemeteries of the site was urn burial (Figure 10.55). The deceased were placed in the bottom of jars measuring 60–90 cm (24–35 in) high and 45–50 cm (18–20 in) in diameter and were buried with few accompanying grave goods. This burial custom was practiced in the Inland Niger Delta into the late nineteenth century. The city reached its height after A . D . 800, when it expanded to 33 ha (80 acres) and a city wall 3.6 m (11.8 ft) wide was constructed around the mound’s 2-km (1.2-mi) circumference. Brick architecture replaced the coursed solid-mud technology that was in use earlier. Closely spaced house foundations of mud bricks are still visible on the surface; mud bricks used in constructing the city wall also are visible today along parts of the mound’s periphery. Furnace parts, forge debris, and crucibles suggest that blacksmiths and coppersmiths worked the iron ore and copper that Jenné-jeno obtained through trade. When Arabs penetrated western Africa at the end of the first millennium A.D., Jenné-jeno and other early West African towns were functioning as fully developed urban centers that were at the apex of regional settlement hierarchies. The rulers of Jenné-jeno presided over a large area of the floodplain extending 60 km (100 mi) downstream. Jenné-jeno began to decline after A.D. 1150 and was abandoned by 1400, as were most of the rural settlements in its hinterland. The reasons for the abandonment are unknown; however, defensive concerns and the spread of
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 495
Figure 10.56 Map of West Africa drawn in A.D. 1375 showing traders on camelback.
Islam after the arrival of the Arab traders are likely factors. The modern city of Jenné (Jenné-jeno means “ancient Jenné”) was established nearby during the twelfth or thirteenth century when Jenné-jeno already was in decline. The new location in an area surrounded by water was easily defended, since it was accessible only by water most of the year. Given the long history of community displacement (both voluntary and coerced) associated with the spread of Islam throughout West Africa, the population base of the new center may have come largely from Jenné-jeno. The new Jenné not only continued to be an important commercial center but also became a center of Islamic learning. The Inland Niger Delta remained an important thoroughfare for exchange throughout much of the second millennium A.D. Another important commercial center on the Niger River was Timbuktu, 500 km (300 mi) north
of Jenné, at the edge of the Sahara. From at least the fourteenth century, the two towns served as major ports of trade in western Africa. North African pottery and blocks of Saharan salt arrived by camel caravan at Timbuktu, where they were shifted to canoes for transport up the Niger River to Jenné (Figure 10.56). At Jenné, they were traded for the rich agricultural produce of Jenné’s hinterland and for gold mined from areas farther south. Historical texts refer to both cities and this “golden trade of the Moors.” Today, modern Jenné is famous for its stunning mud architecture and distinctive mosque that reflects both indigenous and North African influences.
(The Granger Collection, New York.)
_ Site Jenné-jeno
495
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 496
Site
:
Great Zimbabwe
1160 m
1180 m
1180 m
1160 m
1140 m
1120 m
An important trading center in south-central Africa
Hill Ruin
1140 m
1120 m
Maund Ruin
N
1120 m
Great Enclosure
0
300 m
Figure 10.57 The major ruins at Great Zimbabwe.
Relatively egalitarian political formations were more resilient in southern Africa than they were in West Africa. Farming and domesticated animals did not spread into southern Africa until the third century A.D., coincident with the appearance of ironworking and the spread of Bantu-speaking peoples. These herder-cultivators worked metal and made pottery. Before their arrival during the Early Iron Age, southern Africa was occupied by hunter-gatherers whose only tools were of stone.
_ _
496
Chapter 10 States and Empires in Asia and Africa
Peasant farmers first settled in south-central Africa, in what is now Zimbabwe (zim-BOB-way), during the fourth century A.D. Northeast of the Kalahari Desert, Zimbabwe is high plateau country (elevation 1200–1500 m, 4000–5000 ft) bounded by rivers. The gently rolling plains are cool and well watered, covered with savanna woodlands that are free of the disease-carrying tsetse fly, the scourge of equatorial Africa. Mineral deposits are abundant. Iron ores, widespread throughout much of southern and eastern Africa, are present, as are copper and gold. The querns and grindstones found in these early farming villages indicate that grain was grown. Sheep and goats were kept, but hunting continued to provide an important source of meat. People lived in permanent villages, some as large as several hectares, located on open ground with little apparent concern for defense. Their huts were constructed of a wooden framework covered with mud. The presence of slag fragments suggests that ironworking was a common skill. Such small iron artifacts as arrowheads, razors, beads, and rings are found in every village. Copper items were rarer; usually no more than a few beads or small strips have been recovered in any one excavation. Handmade ceramics were produced, of which there were several distinct regional variants. Contact with the east coast of Africa is indicated by the presence of glass trade beads and occasional marine shells. Soon after their introduction, cattle became important both culturally and economically. As in many African societies today, the size of one’s cattle herd was probably a sign of status and a basic means of converting grain surpluses into more permanent kinds of wealth. The advent of cattle herding did not eliminate hunting. Grain
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
crops—particularly sorghum, finger millet, cowpeas, and ground beans— also were cultivated. Bananas were introduced from Indonesia by transIndian traders around the ninth century A.D. During the later prehistory of Africa (in the past thousand years), complex states emerged in the central and southern regions. Groups such as the Karanga were led by powerful chiefs, priests, and traders. They had contacts with societies outside the continent, and at the advent of written history, several of them were still actively involved with foreign merchants. Two of the largest early states in southern Africa were centered on Mapungubwe and Great Zimbabwe (see Figure 10.53). Mapungubwe, the earlier of the two, is located on top of a large sandstone outcrop that rises abruptly from the arid valley of the Limpopo River about 320 km (200 mi) south of Great Zimbabwe. The earliest farmers in the region built a number of villages close to the river after A.D. 800. Their economy was based on cattle, sheep, and goats, and they obtained such goods as cowrie shells and glass beads through Indian Ocean trade. Excavations at one of these early village sites (Bambandyanalo), only a few kilometers from Mapungubwe, revealed a large cattle enclosure, crude beakers and bag-shaped pots, grindstones, and a few iron tools. The first buildings were constructed on Mapungubwe Hill early in the twelfth century A.D. Excavations at the site have revealed a succession of houses and richly adorned burials accompanied by gold beads and bangles. Mapungubwe quickly became one of the largest towns in the region, controlling a hinterland of lesser settlements up to 60 km (37 mi) away. Its inhabitants specialized in various crafts, including working ivory into bracelets, making bone points, and weaving. However, the real base of Mapungubwe’s power came from its intermediary role in coastal trade and the wealth of gold and animal products from its hinterland. When Mapungubwe was at its peak, in the mid-thirteenth century
Page 497
A . D .,
Great Zimbabwe was a smaller district center. Mapungubwe later went into decline just as Great Zimbabwe was reaching its greatest size and influence. The eventual abandonment of Mapungubwe may in part have been due to Great Zimbabwe’s seizure of the gold trade and exchange routes to the coast. Great Zimbabwe, the largest and most famous site of the Karanga, is located in the central region of Zimbabwe, on a tributary that eventually drains into the Indian Ocean (Figure 10.57). The area is composed of granite hills, some of which are enormous, bare, rounded domes. Because of their size, these granite features affect rainfall patterns so that the prevailing southeasterly winds drop more rain here than in neighboring areas. To the north, the site is bounded by a narrow ridge of granite that forms a 91-m (300-ft) cliff, strewn with massive boulders. Just south of the Great Zimbabwe ruins, the land descends into drier, more open grasslands suitable for cattle. Slabs that break off the granite domes provide abundant building material. The Karanga began to build stone structures, including field walls, terraces, and stone enclosures, sometime after A.D. 1000. The first stone structures at Great Zimbabwe, built after A . D . 1250, were placed on top of the
Figure 10.58 The Hill Ruin at Great Zimbabwe consists of a series of stone enclosures built on top of a steep, rocky cliff, possibly for defense. N
Balcony
Eastern Enclosure
Western Enclosure
Cave
Cliff edge Poorly coursed walls 0
Regularly coursed, dressed walls
Uncoursed, ruined, or restored walls
50 m
_ Site Great Zimbabwe
497
pri31057_ch10.pp436-505.qxd
10/1/09
Figure 10.59 The Great Enclosure and other stone ruins at Great Zimbabwe. The Great Enclosure is the largest known prehistoric structure in sub-Saharan Africa. Inside the Great Enclosure were smaller stone structures, which are thought to have housed the site’s ruling families.
_ _
498
8:22 AM
Page 498
high cliff, possibly for defense. Simple stone walls enclosed platforms that held pole-and-mud houses. The walls do not follow an obvious plan. The only openings are narrow doorways, topped with simple stone lintels. The quality of the walls varies, from uncoursed sections of irregularly shaped rocks to coursed walls of granite blocks that were carefully matched. The buildings at the site consist of two groups, one on the steep, rocky cliff and the other on the adjacent valley floor. On the cliff, called the Hill Ruin, well-coursed walls were linked to natural boulders, forming a series of easily defended enclosures (Figure 10.58). The largest and most substantial structure on the hill, called the Western Enclosure, consisted of two curved walls, over 9 m (30 ft) high, circling an area greater than 45 m (150 ft) in diameter. At the other end of the clifftop was a smaller structure, the Eastern Enclosure, bounded by boulders on the north
Chapter 10 States and Empires in Asia and Africa
and a stone wall on the south. Inside this structure were groups of circular stone platforms that held many monoliths. The presence of figurines, including seven carved soapstone birds, suggests that this enclosure was the ceremonial center of the site. The carved birds, about 36 cm (14 in) high, were placed on top of 1-m (3-ft) stone columns. Nothing like these stone carvings has been found elsewhere. In the valley below, larger, freestanding walled enclosures were built surrounding circular pole-and-mud houses. This pattern is especially clear at the Maund Ruin at the edge of the site, where 29 separate stone walls were built. The walls abut ten circular dwelling huts, forming nine separate courtyards, each entered through doorways in the stone walls. These enclosures form single, functional units. Both in the valley and on the hill, large middens of domestic debris accumulated outside most of the enclosures.
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 499
Figure 10.60 The massive outer wall of the Great Enclosure, capped by the chevron pattern.
One enclosure at the opposite end of the valley from the Hill Ruin, the Great Enclosure, was especially large and complex, with a perimeter wall over 10 m (33 ft) high and 5 m (16 ft) thick (Figure 10.59). The outer wall was over 240 m (800 ft) long, forming an irregular ellipse with a diameter of 89 m (292 ft). The top of the wall is decorated with a band of two lines in a chevron pattern (Figure 10.60). There are several entrances into the enclosure on the north and west sides of the wall. Containing more stonework than all the rest of the ruins at Great Zimbabwe combined, this wall is the largest prehistoric structure in subSaharan Africa. Several smaller walled enclosures are situated within this outer wall, containing dwellings that housed the ruler and his family (Figure 10.61). The most striking construction inside the Great Enclosure is a solid circular stone tower rising 10 m (33 ft) from its base, which is 6 m (20 ft) in diameter. Called the Conical Tower, this structure was surrounded by platforms and large monoliths. The function of these monoliths, also associated with the Hill Ruin, remains somewhat of a mystery. Their distribu-
tion was not random; they were placed in areas having a sacred character. The stone enclosures at the core of Great Zimbabwe—covering 40 ha (100 acres)—are the largest and most elaborate of the more than 150 similar stone structures constructed across the high granite region of the Zimbabwe plateau. Many of these sites were small, having between one and five small enclosures surrounded by freestanding walls. The pottery at all these sites was similar to that at Great Zimbabwe. The architectural florescence at Great Zimbabwe was linked to the development of a powerful political authority. The construction of the extensive stone walls clearly required an organized labor effort. Centralized control of expanded trade links with Indian Ocean polities may have been a significant factor, and Great Zimbabwe became an important commercial center, both locally and regionally. Specialized craftworkers made simple forged iron tools, such as hoes, axes, and arrowheads. They alloyed copper with tin and made coiled wire bracelets and pins, needles, and razors, and used imported gold to make bracelets, anklets, and beads. These metals were worked
African cities were in many ways similar to other cities worldwide and were easily recognized as such by Arab and European visitors. —Chapurukha Kusimba (2008)
_ Site Great Zimbabwe
499
pri31057_ch10.pp436-505.qxd
10/15/09
11:06 AM
Page 500
North entrance Northeast entrance
Parallel passage
West entrance
Platform
N Conical Tower
Ch
Regularly coursed, dressed walls
Poorly coursed walls
0
500
att
ern
Uncoursed, ruined, or restored walls
50 m
Figure 10.61 The Great Enclosure, which contains more stonework than all the other ruins at Great Zimbabwe combined.
_ _
p e vro n
on a small scale in certain enclosures set aside for specific tasks. The presence of large numbers of spindle whorls, made from both potsherds and soapstone, indicates that cotton textiles were woven at Great Zimbabwe. The prosperity of Great Zimbabwe was based largely on its monopolization of coastal and long-distance trade, which it had earlier wrested from Mapungubwe. Exchange with Africa’s east coast, which was visited by Arab and Indian merchants, provided a large number of exotic objects, including Persian and Chinese pottery, Near Eastern glass, and cowrie shells from the beaches of the Indian Ocean. These
Chapter 10 States and Empires in Asia and Africa
items were prestige goods, essential to demonstrating rank. Porcelain, glass, and trinkets were found in one hoard at Great Zimbabwe in association with iron gongs, hoes, and seashells. The latter items are still recognized in the region today as symbols of chiefly authority. The traders at Great Zimbabwe also received products from the surrounding area, through either patronage or tribute. Copper was imported from the northern edge of the plateau in the form of standardized ingots. With the initiation of goldworking, the gold trade on the plateau became important by the late twelfth or early thirteenth century A.D. Zimbabwe reached its period of greatest influence between A.D. 1350 and 1450, when the settlement, with a population estimated between 12,000 and 20,000, extended over 700 ha (1700 acres) and controlled an area of approximately 100,000 sq km (38,600 sq mi) between the Zambezi and Limpopo rivers. The balance of power then shifted to a more northerly center on the Zambezi River. The increasing importance of copper from the north may have been a factor in redirecting trade. Later, trade and communication routes, focused on the Zambezi, may have bypassed Great Zimbabwe. The Portuguese penetrated the area in the sixteenth century and established a fort on the east coast. Their attempt to control the gold exchange of south-central Africa further disrupted trade, and the Karanga empire disintegrated by the end of the century. Today, Karanga ruins stand tall in the plateau country of Zimbabwe. The word zimbabwe means “stone houses” or “venerated houses,” and the Karanga used it to identify the houses of chiefs. It is these ruins that now give their name to a contemporary nation.
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 501
Images and Ideas Theories of State Development Changing views on the rise of complex polities and urban societies
T
he early civilizations that arose millennia ago in regions of Asia and Africa all were followed by episodes of political centralization and fragmentation. Over the past century, we have come to document and describe those waves of change more completely. In fact, in all regions of the globe, the emergence of urban societies was followed by episodes of change during which time-specific centers gained and lost in importance and power. Nevertheless, the timing and particular patterns of transition were not always the same. It is important to note that the causal factors that account for those episodes of rise and fall, growth and collapse, remain a matter of debate. In the discussion that follows, we examine some of the ideas and models that have been advanced to account for the emergence of early civilizations. In many respects, the scholarly advances that have taken place over the past decades owe much to the comparative and systematic approaches that were outlined by the prehistorian Vere Gordon Childe. Over a half century ago, V. Gordon Childe described the essence of what we mean by civilization in a list of ten characteristics (even if he did not succeed in providing a scientifically precise or universally acceptable definition). Charles Redman, of Arizona State University, subsequently organized Childe’s indices into a list of primary and secondary characteristics. The primary characteristics are economic, organizational, and demographic in nature and suggest fundamental changes in societal structure. They include (1) cities—dense, nucleated demographic concentrations; (2) full-time labor specialization; (3) state organization, based on territorial residence rather than kin connections; (4) class stratification— the presence of a privileged ruling stratum; and (5) the concentration of surplus. According to Redman, Childe’s secondary characteristics serve to document the existence of the primary criteria. They are (1) monumental public works; (2) longdistance exchange; (3) writing; (4) arithmetic, geometry, and astronomy; and (5) highly developed, standardized artwork. Childe’s criteria, particularly the secondary ones, are certainly not without problems. For example, the Inca, who established the largest pre-Columbian empire in the Americas, did not have a formal system of writing in the traditional sense. Conversely, many societies that are not consensually recognized as civilizations built monumental edifices, engaged in long-distance exchange, crafted wonderful (and reasonably standardized) artwork, and were very aware of astronomical cycles. Even the primary criteria are subject to discussion, since kinship is known to have played a very strong organizational role in both Native American and early Chinese civilizations. Although it is next to impossible using archaeological data to distinguish full-time from part-time craft specialization, occupational specialization (of an unknown degree of intensity) is often found at archaeologically known sites that are not traditionally conceptualized as civilizations. Nevertheless, not only do Childe’s criteria provide a valuable starting point for discussion, but most of them also can be examined archaeologically. To the researcher, they provide a more useful starting point than, for example, the frequently cited definition of the state as an institution that monopolizes force, a characteristic that cannot be subjected to straightforward archaeological investigation. Given the difficulties in defining the state and civilization, as well as the evident variety in human societies and sequences of societal change, it is not surprising that no single, satisfactory explanation has been developed to account for
< ( : < V : y < p : W
The regularities set forth by Steward, and the models constructed by Wittfogel, Carneiro, and others, have guided a great deal of fruitful research leading to knowledge about the material forces important in the rise of complex chiefdoms and states. What constructs, however, will guide us in the future? —Henry T. Wright (1986)
Images and Ideas Theories of State Development
501
_
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
If population pressure was a problem in the Nile Valley, a circumscribed situation par excellence, the Pharaoh should have been delighted to see the last of Moses and his cogeners [followers]. That he was not, and indeed expended considerable effort to retain the Jews within his territories, suggests that population pressure was not perceived as a threat to social stability. —A. Terry Rambo (1991)
Page 502
these transformations. In anthropology, current interpretive perspectives can be subdivided into integrative and coercive models. Integrative approaches emphasize coordination and regulation as roles of emergent institutions. The alternative coercive theories stress the role of the developing state in the resolution of intrasocietal conflicts that emerge from disparities in wealth. These alternative frameworks have their philosophical roots at least as early as the fifth century B.C., when the Greek writer Thucydides described the Peloponnesian War and its combatants. Thucydides compared different organizational frameworks, contrasting the democratic and the oligarchic. The former, typified by Athens under the ruler Pericles, was characterized by government through cooperation, with the populace described as benefitting from state policies and services. Sparta, which typified the latter, more coercive, governing structure, was ruled by a propertied class that controlled decision making to maintain their disproportionate wealth. Most states integrate as well as coerce, although their degree of reliance on different governing strategies certainly can vary. For archaeologists, as well as other social and historical scientists, the decipherment of different organizational strategies is a promising domain for research. Yet in modeling the evolution of early states, researchers should recognize that government strategies can undergo change. For example, institutions may initially develop to serve integrative or regulative roles. Once established, they may become more coercive in the face of new challenges or to maintain whatever benefits their decision makers may have accrued. Therefore, the functions of a governing institution may not provide a complete picture of why that institution arose in the first place. The preceding discussion contrasts the explanatory merits of integrative and coercive frameworks. A second analytical pathway compares the relative utility of different “prime-movers,” key factors that are proposed to account for many, if not all, cases of state development. Karl Wittfogel proposed water control (irrigation) as the key variable in the rise of the “hydraulic state.” Wittfogel saw water as having unique properties, essential for agriculture in the dry lands where many of the world’s early states developed, yet manipulable by people in ways that other environmental resources are not. Nevertheless, although largescale canal irrigation systems were eventually used in the domains of many early states (such as in Mesopotamia), the temporal sequence of state formation and the construction of these grand irrigation networks is not clear. In other areas, such as Mexico’s Valley of Oaxaca, it appears that most pre-Colombian watercontrol devices could have been managed by a few households at most. Recent ethnographic research also indicates that large-scale irrigation networks do not necessarily require centralized administration. A second prime-mover, demographic pressure, places the primary cause of political change on imbalances between a human population and its available food supply. Influenced by the work of the agricultural economist Esther Boserup, proponents of this view have turned the work of Thomas Malthus on its head. In the late eighteenth century, Malthus argued that the advent of agriculture led to the production of food surpluses, thereby making human population growth possible and increasing the availability of leisure time. Yet anthropological work, spearheaded by Robert Carneiro, of the American Museum of Natural History in New York; Marshall Sahlins, of the University of Chicago; and Richard Lee, of the University of Toronto, has questioned the long-held dogma surrounding surplus and leisure time. As Kent Flannery synthesized, The cold ethnographic fact is that the people with the most leisure time are the hunters and gatherers, who also have the lowest productivity; even primitive farmers don’t produce a surplus unless they are forced to, and thus the challenge is getting people to work more, or more people to work. With better technology, people simply work less; what produces surplus is the coercive power of real authority, or the demands of elaborate ritual. (1972a, pp. 405–406)
_ _
502
Chapter 10 States and Empires in Asia and Africa
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 503
Recently, anthropologists have questioned the arguments of Carneiro, Lee, and Sahlins concerning leisure time. They note that most hunter-gatherer populations suffered seasonal or periodic food shortages or frequently lacked certain key resources, such as fat or protein. Yet the fact remains that, except when encouraged, few hunter-gatherers or village people produced a great deal more than their families required. Primarily concerned with the contemporary Third World (where runaway demographic growth is not unusual), Boserup argues that technological changes and increased productivity also could be spurred by excessive population. Archaeological adherents of Boserup’s position view ancient population growth as an independent variable and the principal cause of social and economic transformations. As we saw in many of the site discussions in this chapter and in Chapters 8 and 9, demographic growth often coincided with episodes of great social change, and in many regions, it was an important variable. Yet correlation does not equal causality. What is not clear in most cases is the nature of the interconnections—whether population growth was the cause or the consequence of political and economic transitions. In rural, preindustrial contexts where child labor can be economically valuable, increases in tribute and the labor demands on households (often associated with political development) can spur cycles of demographic growth, as families opt to have more children. In other words, political and economic strategies can greatly influence demographic change. In many of the cases we have examined, the nucleation of population around an emerging center also may have been spurred by in-migration, as people were attracted by or coerced to settle near an increasingly powerful institution. Furthermore, population growth does not necessarily imply population pressure, the latter being a notoriously difficult concept to measure. Archaeological and historical findings from many areas indicate that long-term population change is not regular, uniform, or ever-increasing, making it theoretically problematic to assume continuous and autonomous growth. Finally, in several cases, archaeological findings have shown that regional populations were markedly below any reasonable estimate of available agrarian production at the time of early state development. Exchange also has been advanced as a prime-mover, although, like warfare, it is practically a human universal and therefore too broadly defined to account for the development of the state. Thus the occurrence of exchange is not as evolutionarily significant as are the nature and mode of the transactions, whether they are monopolized or controlled and by whom, the volume of the transactions involved, and the kinds of items moved (and their local importance). Until these considerations are empirically considered and refined theoretically, exchange cannot be convincingly employed as a prime-mover in state development. Another general model for state development was proposed by Carneiro. Stimulated by his ethnographic work among tribal groups in the Amazon Basin of South America, Carneiro has suggested that warfare and bounded conditions may account for the origin of states. Carneiro recognized that warfare, almost universal in human society, cannot alone account for the rise of the state. Warfare was present (even endemic) in many places, yet the state never formed in those areas. In Carneiro’s scenario, a population first must have increased in size beyond the limits of its local resources. Circumscription by either environmental (mountains, oceans, rivers) or social (neighboring groups) boundaries would then require warfare and conquest to obtain more food. Although warfare does appear to have been an important factor in some cases, such as the rise of Monte Albán in the Valley of Oaxaca, Mexico (see Chapter 8, p. 342), Carneiro’s formulation, which tends to make warfare and population pressure into dual primemovers, has not met with unanimous support. For example, it is hard to envision how circumscription played a major role on the extensive, flat North China Plain,
The notion that the study of five or six precocious civilizations would inform us about the factors crucial in the rise of civilization in general is seductive. Throughout the second half of the twentieth century, research focused not only on these earlier developments but also on the supposed “core” areas of these developments, envisioned as the regions in which “breakthroughs” in cultural organization took place and from whence they “diffused” to peripheral regions. As research in the supposed peripheral regions has progressed, we have had some surprises, and empirical knowledge demands that we change our general conceptions. . . . I argue that the idea that civilizations have a single heartland is a product, in part, of the success of elites in particular regions in dominating the historical record and, in part, of the state of archaeology in the twentieth century. —Henry T. Wright (2005)
circumscription The process or act of being enclosed by either environmental boundaries, such as mountains, oceans, and rivers, or social boundaries, such as neighboring groups of people.
Images and Ideas Theories of State Development
503
_
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
I have sometimes compared models to ships. What interests me, once the ship is built, is to launch it, to see if it floats, then to make it sail wherever I please, up and down the currents of time. The moment of shipwreck is always the most meaningful. —Fernand Braudel (1970)
Men make their own history, but they do not make it just as they please; they do not make it under circumstances chosen by themselves, but under circumstances directly encountered, given and transmitted from the past. —Karl Marx (1963 [1852])
w w w. m h h e . c o m / p r i c e i p 6 e
For more review material and study questions, see the self-quizzes on your online learning center.
p
Page 504
where the early Chinese state appears to have arisen. Even in Oaxaca, the connection between the concentration of political power and militarism began during a period when the regional population was small. Today, most archaeologists have adopted multivariate approaches, recognizing that the process of state development was probably triggered by a suite of factors (including some of the prime-movers), rather than a single causal stimulus in each instance, and that even the same set of factors may not have been involved in each case. Our examination of state formation in Asia and Africa illustrates that factors such as population growth, new technologies, changing exchange and interaction (including warfare) patterns, and shifts in the organization of labor and specialization were often intertwined with episodes of managerial restructuring, yet we have sorted out neither the specific interlinkages between those factors nor their relative importance in each case. We also need to work hand-in-hand with natural scientists to document the web of interconnections between humans and their natural environments. Studies in many areas of the globe have now revealed the impacts that humans can have on their immediate settings: flora, fauna, patterns of erosion, desert creation, and more. Such anthropogenic processes and events also require careful analysis before we can understand long-term regional histories. The contingent or historical nature of social change provides a further challenge. In each region, earlier changes always constrain and underpin subsequent shifts. Consequently, the nature of the states in each of these regions varied somewhat because of the nature and organization of the specific polities that preceded them. If we concede that the rise of new forms of government are often accompanied by other significant (and interdependent) shifts at both higher and lower scales (e.g., households, boundary relations), then the analytical tasks in front of us seem all the more challenging. For each region, archaeologists will need to collect information at multiple scales from the activity area to the house, the community, and the region, and even learn to conceptualize at larger scales. All these vantages will have to be examined over time. Nevertheless, we can take heart in the realization that there has been tremendous progress in the study of the state during the past several decades. Recent archaeological surveys, large-scale excavations, the study of ancient households, and ethnohistoric breakthroughs, which have helped us unravel some of the ideological changes that made new managerial formations possible, have enriched the empirical foundation necessary for examining this key societal transformation. In Southwest Asia, South Asia, Egypt, North China, and sub-Saharan Africa, recent findings enable us to refine and improve our models as we expand our knowledge of the history of each of those regions. If these contributions continue apace (especially in the face of our dwindling, threatened archaeological record), and if a series of crucial definitional and theoretical challenges are met, the opportunity for taking giant steps forward in our understanding lies immediately ahead.
DISCUSSION QUESTIONS 1. How did the development and the role of writing vary in the early civilizations of Mesopotamia, the Indus Valley, Egypt, China, and Mesoamerica?
anthropogenic The term used to
describe an effect or process resulting from human activity. The creation of pasture from forest through intentional burning is an example. At times, human abandonment of a site or area (as well as initial occupation) also can set off environmental changes.
_ _
504
2. Was the development of irrigation systems a factor in the rise of most early complex societies? Why do you think so? If not, what other factors were important? 3. The early Egyptian state appears to have been more centralized than the Indus civilization. Can you offer some thoughts as to why? 4. Contrast the nature of rulership in two of the early civilizations discussed in this chapter. Can you account for or explain these differences?
Chapter 10 States and Empires in Asia and Africa
pri31057_ch10.pp436-505.qxd
10/1/09
8:22 AM
Page 505
5. In sub-Saharan Africa and Southeast Asia, external contacts and exchange are thought to have had a role in the rise of early civilizations. How did such exogenous relations interplay with local traditions and developments in each region?
z
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e.
Alcock, S. E., T. N. D’Altroy, K. D. Morrison, and C. M. Sinopoli, eds. 2001. Empires: Perspectives from archaeology and history. Cambridge: Cambridge University Press. An edited collection that brings a comparative and global perspective to historically known empires. Carneiro, R. L. 2003. Evolutionism in cultural anthropology: A critical history. Boulder, CO: Westview. Traces the interaction of evolutionary thought and anthropological theory from the nineteenth to the twenty-first century. Feinman, G. M., and J. Marcus, eds. 1998. Archaic states. Santa Fe, NM: School of American Research Press. A recent collection of essays that highlights the diversity of ancient states. Kenoyer, J. M. 1998. Ancient cities of the Indus Valley civilization. Oxford: Oxford University Press. A timely overview by an area specialist. Murowchick, R. E., ed. 1994. Cradles of civilization: China. Norman: University of Oklahoma Press. A wellillustrated collection of articles examining the early civilizations of China. Rothman, M. S., ed. 2001. Uruk Mesopotamia and its neighbors: Cross-cultural interactions in the era of state forma-
tion. Santa Fe, NM: School of American Research. A set of scholarly essays that presents up-to-date information on this early civilization. Smith, M. L., ed. 2003. The social construction of ancient cities. Washington, DC: Smithsonian Institution Press. A collection of papers that provides a bottom-up perspective on ancient cities in several world areas. Trigger, B. G. 2003. Understanding early civilizations: A comparative study. Cambridge: Cambridge University Press. A detailed comparative study of the seven bestdocumented early civilizations in the world. Wright, H. T. 1986. The evolution of civilizations. In American archaeology, past and future, ed. D. J. Meltzer, D. D. Fowler, and J. A. Sabloff. Washington, DC: Smithsonian Institution Press. An analytical synthesis of long-term change in several world regions. Yoffee, N., and G. L. Cowgill, eds. 1988. The collapse of ancient states and civilizations. Tucson: University of Arizona Press. A comparative collection of scholarly papers on collapse by experts from several disciplines.
_ Suggested Readings
505
pri31057_ch11.pp506-559.qxd
10/1/09
C
12:12 PM
H
Page 506
11
A
P
T
E
R
Figure 11.1 Gold death mask from a shaft grave at Mycenae, Greece, ca. 1550 B.C.
_ _
1
1
pri31057_ch11.pp506-559.qxd
10/1/09
12:12 PM
Page 507
< Prehistoric Europe( Introduction : < T V : y < p : W From the First Farmers to the Roman Empire A story of change
he funeral mask of “Agamemnon” from the shaft graves at Mycenae, Greece (Figure 11.1), captures several aspects of the later prehistory of Europe. It depicts a warrior-king (although almost certainly not Agamemnon) who ruled an agricultural society. It is made of an exotic raw material, gold, traded longdistance, probably from Bulgaria. From the appearance of the first farmers to the historic events of the Roman conquest, the archaeology of the past 9000 years in Europe is a story of the introduction of agriculture, of technological innovation and the use of metals, of the growth of regionalism and warfare, and of the development of economically and politically powerful groups. This period is of interest today because many of the basic tenets of Western civilization come from prehistoric Europe. Languages, customs, traditions, forms of government, and many of our fundamental solutions to the uncertainties of the world emerged in Europe during this period. The period also interests us because, in the span of a few thousand years, we can trace the development of human societies from small, simple bands of hunter-gatherers to large, complex states with thousands of citizens. European archaeology is also fascinating because a long history of research has produced a number of remarkable finds. It is useful to consider the prehistory of Europe in millennia, beginning with the arrival of agriculture before 7000 B.C. and ending with the decline of the Vikings around A.D. 1000. The millennia B.C. run backward; the first millennium B.C. is 1000–0 B.C., and the eighth millennium is 8000–7000 B.C. Dates designated A.D. run forward. We are now in the first years of the third millennium A.D. Eighth Millennium (8000–7000 B.C.) Excavations at the site of Franchthi Cave in Greece document the introduction of agriculture to southeastern Europe shortly before 7000 B.C. Many major innovations—domesticated plants and animals, pottery, iron, writing, and others—came to Europe from Southwest Asia, appearing initially in the southeast and moving gradually to the north and west (Figure 11.2). Incipient farming communities appeared in the Aegean area and Greece shortly before 7000 B.C. and arrived in northwestern Europe around 4000 B.C. Seventh Millennium (7000–6000 B.C.) During this millennium, farming and pottery followed two main pathways from Greece into Europe: by land into the center of the continent, and by sea along the north coast of the Mediterranean. This package of pottery, mud-brick houses, and domestic plants and animals had spread by 6500 B.C., via the Danube and other rivers, to the Balkan Peninsula of southeastern Europe (Bulgaria, Romania, Hungary, and the former Yugoslavia). Along the second path, distinctive Cardial pottery, decorated with the scalloped edge of a seashell, wheat, and the bones of domesticated sheep appeared in caves and rockshelters along the Mediterranean shore after 6000 B.C. The coastal location
507
CHAPTER OUTLINE
Introduction From the First Farmers to the Roman Empire 507 SITE Franchthi Cave SITE Varna
511
513
CONCEPT The Iceman SITE Charavines
516
520
CONCEPT The Megaliths of Western Europe 524 SITE Stonehenge
528
CONCEPT The Aegean Bronze Age 532 SITE Knossos
SITE Mycenae
535
538
CONCEPT The Bronze Age North of the Alps 542 SITE Borum Eshøj SITE Vix
544
547
CONCEPT The Bog People SITE Maiden Castle
550
552
Images and Ideas
Lessons from Prehistoric Europe 555
_
pri31057_ch11.pp506-559.qxd
10/1/09
12:12 PM
Figure 11.2 The spread of agriculture from Southwest Asia into Europe in time and space. The map shows the approximate dates for the major stages in the expansion of farming across Europe.
Page 508
Limit of agriculture
7000 B.C. 6000 B.C. 5500 B.C. After 5500 B.C.
0
100
0 100
w w w. m h h e . c o m / p r i c e i p 6 e
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
y
Was the spread of agriculture into Europe slow or fast? Gradual or jumpy?
millennium A period of 1000 years. lake dwelling The former name for
Early Neolithic lakeshore settlements originally thought to have been built over the water. Bandkeramik An archaeological cul-
_ _
ture of the early Neolithic in central Europe, referring to the style of pottery: linear bands of incised designs on hemispherical bowls.
508
300 Miles 300 Kilometers
and sporadic distribution of these sites suggest a spread of agriculture arriving by sea. Sixth Millennium (6000–5000 B.C.) This time witnessed the flowering of Neolithic cultures in southeastern Europe—the appearance of large settlements, elaborate religious systems, copper mining, and extensive trade networks. A number of the many large tells in the region, such as Starcevo outside Belgrade in Serbia and Karanovo in Bulgaria, have been excavated and provide a sequence of these events. These sites contain houses, elaborate pottery, and copper artifacts. Copper ore was mined from rich deposits in Serbia and elsewhere, providing raw material for the oldest metal production in the world. Finished axes and copper jewelry were traded throughout Europe. The cemetery at Varna documents the extraordinary wealth in copper and especially gold that accumulated in southeastern Europe during this period. During this time, agriculture continued to spread across Europe. The coastal branch moved inland from the Mediterranean shore into Spain, France, and Italy. By 4000 B.C., farming communities were distributed across most of southern Europe. The Iceman came from one of these communities in northern Italy before he died at the top of the Alps. A number of very well preserved settlements, often described as lake dwellings, are known from Switzerland, France, and other countries around the Alps. As one example, the village of Charavines in southeastern France provides an extraordinary glimpse into the Neolithic because of the remarkable preservation of wood and other organic materials at the site. The interior arm of agricultural expansion reached into central Europe during this same period. The early Neolithic communities in this area belong to the Bandkeramik, or Linear Pottery, culture. The Bandkeramik had its origin around 5700 B.C. in villages along the middle Danube and its tributaries in eastern Hungary. From this core area, farmers migrated east, north, and west, across the loesscovered valleys of central Europe to Belgium, southern Poland (Figure 11.3), and Ukraine. Within a period of a few hundred years, small farming villages appeared throughout this area. Both the rapid expansion of the Bandkeramik and the uniformity of its architecture, artifacts, burials, and settlement plan are remarkable.
Chapter 11 Prehistoric Europe
pri31057_ch11.pp506-559.qxd
10/1/09
12:12 PM
Page 509
Fifth Millennium (5000–4000 B.C.) This was a time of expansion and consolidation as Neolithic groups filled in much of southern and central Europe and regional differences emerged in ceramic styles. Neolithic groups began to develop local traditions as independent societies. Fourth Millennium (4000–3000 B.C.) The beginning of this period witnessed the last stage in the spread of agriculture to northwestern Europe—the Netherlands, northern Germany, northern Poland, southern Scandinavia, and the British Isles. Neolithic farming societies in western Europe quickly began to erect megalithic structures—large stone tombs and monuments such as Stonehenge. Across Europe, in spite of such widespread traditions and growing trade, conflict and warfare are documented in the increasingly defensive nature of settlement location. Third Millennium (3000–2000 B.C.) Major innovations appear, including the introduction of bronze, new weapons, the wheel, draft animals and the plow and oxcart, the horse and chariot, and extensive maritime contacts. Bronze first appeared in the form of weapons and jewelry and was largely in the possession of the elite. Bronze swords and spears must have provided a distinct military advantage. Trade and conflict appear to have escalated during this period as well. Second Millennium (2000–1000 B.C.) The second millennium B.C. was important in the Aegean area. First on Crete, and later in Greece, Bronze Age lords directed powerful polities. The palaces of Knossos and Mycenae provide evidence of the vitality of these early states. The palace economy involved writing systems, craft specialization, taxation, and extensive trade networks. This commerce attracted goods and materials from most of temperate Europe to the Aegean. The second millennium B.C. was also the time of the Bronze Age north of the Alps. Large polities arose in those areas where raw materials and trade routes coincided. The elaborate tombs of elite individuals under large earthen mounds in southern England, the Czech Republic, Spain, and southern Scandinavia are evidence of local wealth and high status in these areas. The Danish site of Borum Eshøj provides examples of such tombs and the monuments of these societies. First Millennium (1000–0 B.C.) This time is known as the Early Iron Age because of the introduction of this new metal. The classical civilizations of Greece and Rome arose in the Mediterranean region during this time. These literate polities with written histories are the subject of classical archaeology. North of the Alps, prehistory continued. In temperate western Europe, Celtic and Germanic tribes with distinctive traditions and art styles were present during the Iron Age. The fortresses of these warrior societies—known by their Roman name, oppida, centers of trade and warfare—dotted the landscape of western Europe. The princess burial at Vix in France documents the wealth and interaction of those Celtic elite; her tomb is located at the foot of Mont Lassois, one of the largest fortified towns of Celtic Europe. Iron Age sacrifices and executions preserved in the bogs of the north provide a startling glimpse of some of the inhabitants of that area. The discipline and strength of the Roman legions finally overwhelmed the Celts, or Gauls, as Julius Caesar and his successors carried the legacy of Rome to western Europe. The Iron Age fortress of Maiden Castle in southern England was razed in A.D. 43 (Figure 11.4). By A.D. 125, the Romans had extended Hadrian’s Wall across the northern border of England. Only those areas north and east of the Rhine River remained free of Roman rule and continued a tribal way of life for a short while longer. Eventually, in the first millennium A.D., the spread of Christianity brought a new religion, and the literate priesthood recorded history, closing the prehistoric part of Europe’s past. The Vikings succumbed to history and Christianity around A.D. 1000.
Figure 11.3 A modern reconstruction of the interior of and the entrance to the Iron Age settlement at Biskupin, Poland. This lakeside site has a long history of occupation from the Bronze Age and the Iron Age.
oppidum A massive fortification in western Europe, often on a hilltop or a bluff, built for defensive purposes during the Iron Age.
Introduction From the First Farmers to the Roman Empire
509
_
pri31057_ch11.pp506-559.qxd
10/1/09
12:12 PM
Page 510
Figure 11.4 Location of and timeline for sites in prehistoric Europe.
Borum Eshøj
Stonehenge
Maiden Castle
Vix Iceman Charavines
Varna
Mycenae Franchthi Cave Knossos
Franchthi Cave Varna Iceman Charavines Stonehenge Knossos Mycenae Borum Eshøj Vix Maiden Castle
8000
6000
_ _
4000
2000 Years
510
Chapter 11 Prehistoric Europe
B.C./A.D.
2000
pri31057_ch11.pp506-559.qxd
10/1/09
12:12 PM
Page 511
Site
:
Franchthi Cave The arrival of the Neolithic
Franchthi (FRAUNK-tee) Cave in southern Greece provides early evidence for the introduction of agriculture to the European continent (Figure 11.5). Today, the cave is situated at the Aegean coast; in the early Holocene, however, the sea was some distance from the cave, and a fairly level plain stood where its waters are now. Investigations at the cave led by Thomas Jacobsen, of Indiana University, revealed layers of trash and other debris that contained a detailed picture of human residence there over the past 20,000 years (Figure 11.6). The early Mesolithic inhabitants of Franchthi exploited a wide range of terrestrial and marine resources at the end of the Pleistocene. Evidence for the hunting of red deer and other large game animals, the collecting of marine mollusks and land snails, and the use of several types of wild plants, including leafy greens, pistachio nuts, almonds, and various grasses, is found in the Upper Paleolithic deposits. The variety and seasonal availability of the foods at the site suggest that the cave may have been occupied year-round by 10,000 years ago. The Mesolithic levels at the site indicate an increasing reliance on the sea. The remains of tuna and the presence of obsidian document the seafaring abilities of the residents around 10,000 years ago. Bones from large tuna constitute about 50% of the animal remains in these cave layers. Measuring up to 2.5 m (8 ft) long and weighing up to 200 kg (450 lb), tuna live only in the deeper waters of the Aegean. The obsidian in some of the artifacts at Franchthi comes from the island of Melos, some 150 km (100 mi) away. After 7000 B.C., domesticated plants and animals were present at Franchthi. Sheep and goats were abundant in the cave levels from this period, as well as domesticated wheats and barley. Red
deer and tuna were less common among the animal remains. Plant species consumed earlier declined, including wild oats, pistachios, and barley. Some stone blades have sickle polish along their edges, a substance that remains when blades are used to cut the stems of plants such as grasses and reeds. The presence of sickle polish indicates that the inhabitants of the cave were harvesting plants in substantial quantities. Coriander seeds were also found among the plant remains. Coriander grows as a weed in cultivated or disturbed areas; its presence suggests that crops may have been cultivated. Finally, the lentils preserved from this level at Franchthi are large, the same size as the later domesticated variety. The occupation area increased in size at this time. A terrace in front of the cave was used for the houses of a
Figure 11.5 Franchthi Cave, on the coast of the Aegean in southern Greece. During the first part of the occupation here, the Mediterranean was some distance from the cave, but rising sea levels gradually brought it closer.
sickle polish A clear polish that forms along the edges of flakes and blades that are used to cut reeds, grass, wheat, and other long-stemmed plants.
_ Site Franchthi Cave
511
pri31057_ch11.pp506-559.qxd
10/1/09
12:12 PM
Page 512
TABLE 11.1 The Major Layers at Franchthi Cave
Period
Sea Level
Proportion of Wild Animal Bones
Late Mesolithic
–25 m
Deer, pig
Middle Mesolithic Early Mesolithic
Deer, pig Deer, horse, goat, pig
Fish
Plants
Years B.P.
Large and common Small and rare Small and rare
Wild barley, vetch, lentils
8500
Wild barley, vetch, lentils Wild barley, vetch, lentils
10,500 10,500
Hiatus Late Paleolithic Upper Paleolithic
–50 m –90 m
Deer, horse, cattle, goat Deer, horse
small farming community with a population of perhaps 75 people. Contacts with other areas increased. New exotic materials such as marble appeared in the cave deposits for the first time. However, the introduction of domestic plants and animals, as well as pottery, Figure 11.6 A view down the “chimney” of Franchthi Cave to the excavations beneath. This large cave contains deposits from the Paleolithic, Mesolithic, and Neolithic of Greece.
_ _
512
Chapter 11 Prehistoric Europe
— —
Wild barley, vetch, lentils —
12,500 22,000
did not disrupt all aspects of Franchthi Cave; there was some continuity with the previous way of life. The domesticated plants and animals appeared rather suddenly at Franchthi and must have come from Southwest Asia. There are several possible scenarios for this introduction of agriculture. One involves the influx of people bringing domesticates to Greece and the Aegean islands. Another possibility is the adoption of exotic domesticates and pottery by the indigenous peoples of southeastern Europe. It seems unlikely that the inhabitants of the cave would have changed completely with the arrival of the Neolithic. Perhaps the domesticated plants and animals were introduced by seafarers from Turkey or Southwest Asia looking for exotic materials to trade or exchange. Whatever the initial mechanism of introduction, these innovations initiated a series of changes that would eventually be felt throughout Europe. Within 3000 years of their first appearance in Greece, farming societies had replaced hunter-gatherers across most of the European continent.
pri31057_ch11.pp506-559.qxd
10/1/09
12:12 PM
Page 513
Site
:
Varna Golden burials on the Black Sea coast of Bulgaria
In 1972, workers digging near the city of Varna on the Black Sea coast of Bulgaria uncovered several graves containing unusual metal objects. Thinking the metal was copper, they called in local archaeologists, who excavated the cemetery from 1973 to 1982. The discoveries at Varna have dramatically altered our impressions of early farmers in southeastern Europe because of the rare and valuable materials that accompanied the burials (Figure 11.7). The discovery and use of metals in the Old World was a relatively slow process. A few small pieces of copper, in the form of jewelry, appeared in Southwest Asia by 7000 B.C. at early Neolithic sites. This was native copper, simply hammered from its original shape into a new form. The melting and casting of copper began in southeastern Europe and Southwest Asia shortly after 5000 B.C. Copper mines were opened in Yugoslavia, and various copper artifacts, primarily axes and jewelry, found their way throughout much of Europe. Gold objects also began to appear during the fifth millennium B.C. The cemetery contained at least 190 graves in an area of 6500 sq m (1.6 acres). The graves are simple rectangular pits with rounded corners, dug into the earth to varying depths, up to 3 m (almost 10 ft). Red ochre was spread over most of the burials. The “copper” the workers discovered was, in fact, gold, and more than 6 kilos (about 14 lb) were found during the excavation of the cemetery. A wide range of other materials were uncovered in the graves as well, including flint, obsidian, bone, clay, ochre, shell, graphite, marble, and copper. Impressions of decayed textiles could be seen on the walls of the graves and were preserved with
Figure 11.7 One of the richest graves at Varna. Note the gold discs around the head and on the chest, the arm rings, the axes, and especially the gold-handled stone axe in the man’s hand.
_ Site Varna
513
pri31057_ch11.pp506-559.qxd
10/1/09
12:12 PM
Page 514
Figure 11.8 One example of the earliest use of gold for decorating pottery at Varna. The high temperatures involved in applying the gold to the ceramics imply sophisticated pyrotechnology.
_ _
cenotaph A grave that does not con-
tain a skeleton.
514
some of the copper artifacts. Radiocarbon methods date the site to approximately 4500 B.C., contemporary with the Vinc˘a culture in Yugoslavia. The dead were buried with their heads toward the Black Sea. Most of the burials were males, lying extended on their backs. A few women were buried in a flexed position on their right sides. There were no graves with children; the youngest individual in the cemetery was about 14 years old. The graves contained relatively few furnishings. Most male burials included a copper axe or pottery and flint tools; female graves held only a few ceramic pots. Almost 20% of the graves did not contain a body—a pattern seen elsewhere in the Neolithic of southeastern Europe. Most of these empty graves, or cenotaphs, contained only a few offerings, usually gold rings and copper axes. Several of the graves were quite distinctive. Three empty graves contained life-size clay masks of human faces. Certain facial features, such as the teeth and decorations on the ear, chin, or forehead, were made from gold plaques. In addition, thousands of gold pins and beads of Spondylus
Chapter 11 Prehistoric Europe
and Dentalium shell, copper needles, marble drinking cups, figurines, and graphite-and-gold-painted pottery were found in these three “mask” graves. The graphite and gold ceramics are exceptional examples of the technological sophistication of the potters of this period (Figure 11.8). Graphite is applied as a powder to the burnished clay surface of the pottery and must be fired at a temperature of at least 1000°C (1800°F) to fix the graphite. Gold powder may have been applied to pottery in a similar fashion, but the exact technique is unknown. Three other empty graves, numbered 1, 4, and 36 during the excavations, contained even more wealth. The offerings and grave goods were arranged in these tombs as though a body were present. Grave 1 contained 1225 gold objects weighing 2093 g (4.5 lb), in 15 groups. (At the price of gold, the value would be roughly $1 million today.) These items included two gold tubes and parts of the shafts of axes. Golden masks were found in several of the graves. Grave 36 included a solid gold axe and shaft, two gold bull effigies, 30 miniature golden horns, and
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 515
Figure 11.9 The contents of one grave at Varna include enormous flint blades, shell beads, gold plaques and rings, gold handles on a shaft, gold armbands, bracelets, and beads, and several copper axes and other tools.
various gold adornments and jewelry, a marble dish, four pottery vessels, and miscellaneous bone and flint tools. Grave 4 contained 320 gold pieces, weighing 1500 g (3.3 lb). The list of items in the grave is remarkable: A stone axe of beautiful workmanship with a tubular gold shaft was placed as if at the right shoulder of the missing skeleton. Other grave goods included: a copper pickaxe of the same shape as found in the . . . copper [mines]; a shafthole axe; a flat axe; a chisel and awls [one with a preserved bone handle]; a dark green stone axe; an enormous flint blade [over 40 cm long]; oblong and rectangular breastplates of gold, having two small perforations in each corner; a circular convex gold disc 7 cm across, lying at the shoulder next to the golden shaft of the axe; one very large globular gold bead placed at the head; round gold earrings; a necklace of annular gold beads; three massive armrings of gold in the center of the grave; 41 circular convex discs of gold with perforations on the side, probably garment ornaments; and a mass of beads of Spondylus and Dentalium shell and of semiprecious orange, red, and black stone. At the head were a large gold-
painted dish and vase and three other pots with lids. (Gimbutas, 1977, p. 48) The graves at Varna provide spectacular evidence for status differentiation in the Neolithic of southeastern Europe. Different categories of people in this society were distinguished by the wealth that accompanied the burials. The rich graves may have been those of religious or political leaders or merchants. Certainly, the location of this important cemetery points to the role of trade and exchange. Varna is located at the shore of a former inlet from the Black Sea, perhaps a natural harbor, and near the mouth of the Danube River. Trade routes for gold, copper, obsidian, marble, shell, pottery, and many other items must have run through or very near this area. Such a strategic location may well have led to the rise of an elite at Varna and the accumulation of extraordinary amounts of wealth. The fact that the wealth of the living was buried with the dead is both a remarkable testimony to the complexity of Neolithic society in the Balkan Peninsula and a boon to archaeology (Figure 11.9).
status differentiation Inequality in human society in which certain individuals or groups have access to more resources, power, and roles than others.
Site Varna
515
_
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 516
Concept
<
The Iceman A frozen mummy from the Neolithic One of the most remarkable finds in the last century was made in 1991 in the high Alps, on the border between Italy and Austria. Hikers noticed a human body, half-frozen, face down, in the snow and ice along the trail (Figure 11.10). The hikers contacted the authorities, who assumed it was the body of an unfortunate mountaineer; a number of people die in the Alps each year. The remains were somewhat carelessly removed from the ice and taken to a morgue. After inspection of the body, and particularly the items found with the corpse, however, it became clear that this was not the result of a skiing accident—it was the frozen mummy of a man from the Stone Age. The body and other artifacts are the highest archaeological finds in Europe, at 3200 m (10,500 ft). The body was taken to the University of Innsbruck for study; a wide range of scientists have been involved in the re-
Figure 11.10 The body of the Iceman as he was first seen in the Italian Alps. This Neolithic man was discovered by hikers during a warm summer when the ice that had covered him for 6000 years melted and exposed his corpse. He was first thought to be an unfortunate climber or skier until the true antiquity of the body was determined.
_ _
516
Chapter 11 Prehistoric Europe
search. More than 150 specialists have been examining all aspects of Ötzi the Iceman, as he has come to be known. The Iceman was approximately 46 years old at the time of his death. He was not a tall man, 159 cm (5 ft, 2.5 in) in height. Sometime fairly soon after his death, snow buried his body; there are no traces of scavenging by birds or other animals. He was mummified, dried by the sun, wind, and ice of the mountaintops. It does appear that during warmer episodes in the past, the body of the Iceman and some of his equipment may have floated in a small lake formed at this place by melting ice. Over time, a small glacier on the mountaintop expanded to cover the depression in which the Iceman lay. Amazingly, this thick, heavy layer of ice did not crush the body, but sealed it into a small depression for 6000 years, as if in a huge freezer. The glacier retreated during the warm summer of 1991 and exposed the corpse.
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 517
Figure 11.11 The body of the Iceman and some of his tools. The artifacts found with the Iceman included his axe, his bow and arrows, some jewelry, mushrooms, and clothing.
The extent of preservation of the body is remarkable; most of the internal organs, as well as the eyeballs, are intact. Although the body was hairless from the effects of freezing and thawing, hair was found around the body. Tattoos were clearly visible on his back and right leg. X rays revealed several broken ribs and indicate that the Iceman suffered from arthritis in his neck, lower back, and right hip. An absence of stomach contents and the presence of material in the large intestine indicate that he had not eaten for 8 hours. His last meal included unleavened bread, some greens, and meat. DNA analysis identified the meat as coming from red deer (European elk). Analysis of pollen in the stomach contents indicates he died between March and June. The Iceman was approximately 46 years old at the time of his death. Radiocarbon dates from the body and the equipment the Iceman carried indicate an age of around 4300 B.C., clearly in the Neolithic period. The Iceman was carrying a substantial amount of gear with him, consisting of seven articles of clothing and twenty items of equipment (Figure 11.11). The items were a bow and quiver of arrows, bowstrings, bone points, a needle, a hafted copper axe, a wooden rucksack frame, two birchbark containers, a hafted knife of flint and its sheath, several flint tools (including a scraper,
an awl, a flake, and a tool for pressureflaking flint), a net (perhaps for catching birds), a piece of ibex horn, a marble pendant, and birch fungus (possibly used as medicine). Most of the arrows were unfinished; he carried several items that were incomplete or in need of repair. No food was found among his possessions. His clothing included a large belt with pouch, holding up a leather loincloth and skin leggings, a coat of deerskin, a cape of woven grass, a conical leather cap with fur on the inside and a chin strap, and shoes of calfskin filled with grass (Figure 11.12). One of the more interesting finds is the copper axe; it is almost pure. This new metal documents the widespread use of copper during the latter half of the Neolithic period. Copper was being mined and smelted in several areas of eastern Europe by this time and traded as far away as Scandinavia. The more difficult questions about the Iceman include where he came from and how he died. He probably came from valleys to the south in Italy, less than a day’s walk away. He carried a small ember of charcoal, for use in starting fires, that came from the wood of trees that grow south of the Alps. Pollen found in his intestines came from the hornbeam tree, which grew only to the south of the Alps. In addition to the pollen, tiny fragments of moss from the species known
Figure 11.12 The Iceman’s outerwear, made of grass and leather. Concept The Iceman
517
_
10/1/09
12:40 PM
Page 518
nt
er
ta
l
pri31057_ch11.pp506-559.qxd
Ve
Vent
GERMANY Innsbruck
ta
l
AUSTRIA
ed
er
Bolzano
Ni
AREA OF DETAIL I TA LY
Ö T Z TA L A L P S
Lake Garda Venice
Verona
Hauslabjoch Ötzi Site
Po River
AUSTRIA
Finailspitze T i s e n ta l (Valdi Tisa)
Bo rd er
I TA LY
Sc (Va h n lS a e
l ta l s les) na
Vernagt Reservoir (Vernago)
Possible route of Iceman Elevation Profile
4,000
Finailspitze 3,514 meters Ötzi Site Tise
ntal
3,000 2,500
Niadertal
Vent
2,000 1,500
Juval Castle
l ista
Ventertal
Vernagt Reservoir
na
1,000
Sch
500
Vinschgau
I TA LY
AUSTRIA
0 0
5
10 15 20 25 Distance (kilometers)
30
Neckera complanta moss sites
au ) h g sta s c no in e V al V (V
Height (meters)
3,500
Juval Castle
35
Figure 11.13 The location of the find of the Iceman and relevant geographic areas. Several lines of evidence suggest that he came from the south into the high Alps.
_ _
From James H. Dickson, Klaus Oeggl, and Linda L. Handley. 2003. The Iceman Reconsidered. Scientific American.
518
Chapter 11 Prehistoric Europe
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
as Neckera complanata were found in his intestines, perhaps used to cover and protect the food that he carried and ate. This species of moss grows only on the southern slopes of the Alps and also indicates that region of Italy as the place from where he came (Figure 11.13). Another line of evidence comes from the isotopes in his teeth, which provide a record of the place of his birth. These isotopes originate in rocks and soil and differ from place to place. The isotopes get into our teeth through the foods we eat. Tooth enamel forms during childhood and does not change during our adult lives. Isotopes of strontium in the food that the Iceman ate as a young child, found in his tooth enamel, also point to a location to the south in Italy as his place of origin. Until recently, there were several theories about how the Iceman died in the high Alps, involving either an accident or some kind of escape. Some suggested he may have been a shepherd in the mountains, caught by an early fall blizzard. However, more re-
Page 519
cent study of the body has revealed several wounds. Deep cuts to his hand and wrist suggest he was in an armed struggle, and an arrowhead lodged in his back may well have been the cause of death. It now seems that the Iceman died as a result of violent conflict. The Iceman is one of those rare finds in archaeology that attract enormous public attention because of the unusual conditions of preservation. The mysteries of the Iceman—how the body was preserved for so long, and how and why he died—add to the aura of intrigue surrounding the discovery. Equally important, however, is how much we are learning from the Iceman about the artifacts, clothing, and equipment of the Neolithic—and, perhaps, especially how those people were not very different from us. The Iceman has now been returned to the cold. His body today is displayed through a small window in a large freezer in a $1.9 million exhibit in a museum in northern Italy.
_ Concept The Iceman
519
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 520
Site
:
Charavines A Neolithic lakeside village The past is also found underwater. Underwater archaeology has been going on for some time. Although survey and excavation conditions are more difficult, the exceptional preservation of organic materials often makes the effort worthwhile. Perhaps the most famous finds have been of Greek and Roman ships in the Mediterranean, but earlier Bronze Age shipwrecks are also well known. In fact, there have been underwater finds from every period of European prehistory. The Upper Paleolithic cave of Cosquer off the southern coast of France is one example. Numerous Mesolithic sites have been found off the coasts of Denmark and Sweden in the Baltic Sea. Several submerged Neolithic settlements have been discovered and excavated in Switzerland and France. Originally described as lake dwellings, these settlements were thought to have stood over the water on raised pilings. It is clear today, however, that the lake lev-
Figure 11.14 Preserved wooden posts on the lake floor at Charavines. The posts were exposed during low lake levels in 1906, revealing the site for the first time. The location of the posts in the modern lake gave rise to the idea that prehistoric peoples lived over the water.
_ _
520
Chapter 11 Prehistoric Europe
els are higher now than in the Neolithic, and these villages, once along the shoreline, were submerged after their abandonment. One of the best examples of such a Neolithic lakeshore settlement is the site of Charavines (SHAR-ah-vans), located between Lyon and Grenoble in southeastern France. The site was originally noticed in 1906 when the tops of hundreds of large wooden posts were seen under the water, approximately 100 m (330 ft) off the shore of Lake Paladru (Figure 11.14). Charavines was excavated by a team from the Institute for Alpine Prehistory at the University of Grenoble, headed by Amié Bocquet. These excavations, under 2–4 m (7–13 ft) of water, lasted from 1972 through 1986 and produced a remarkable array of food remains and items of wood, bone, fiber, and other materials, preserved in the oxygen-deficient mud of the lake bottom (Figure 11.15). The finds from
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 521
Figure 11.15 Artist’s interpretation of a diver at work. Underwater archaeology takes place all over the world, exploring a variety of time periods.
TABLE 11.2 The Distribution of Animals during the Later Occupation at Charavines, Based on 504 Individuals Percentage of Individuals Red deer Wild boar Goat Sheep Roe deer Cattle
Percentage of Meat Used*
60.5 22.7 6.0 5.0 4.5 1.3
61.0 29.3 1.7 1.7 1.3 5.0
Wild or Domestic W W D D W D
*Estimated meat obtained, assuming all animals were eaten. Centre de documentation de la Préhistoire Alpine 53 rue du Drac, 38000 Grenoble (France) Tél/Fax: (33) 04 76 96 34 24 Pré
[email protected] http://perso.wanadoo.fr/..ctredocumprehistalpine/cha8an.htm
this site prompted Bocquet to humorously suggest changing the other name for the Neolithic from the Stone Age to the “wood age.” Hundreds of specialists have worked for years analyzing the various materials recovered in the excavations. The excavations at Charavines revealed that there had been two major phases of occupation along this lakeshore more than 5000 years ago. Each one lasted 20–25 years, separated by about 3 years of abandonment. This detailed information is based on dendro-
chronological analysis of the wooden beams used in house construction. Most of the beams have traces of bark remaining, so the date when the trees were felled can be determined from the state of growth of the tree rings. Fir trees were felled during winter for timber to build the first two houses at Charavines; 9 years later, new posts were cut and used to rebuild the houses, and 8 years later they were rebuilt again. The number of structures per settlement phase ranged from three to eight. The excavators were able to map
_ Site Charavines
521
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Figure 11.16 Artist’s interpretation of the Neolithic village at Charavines around 2600 B.C. The age has been determined by the dendrochronology of the wooden construction posts. The dendrochronology permits very detailed determination of the year of the construction of the houses.
_ _
522
Page 522
the precise locations of the houses. The house contents were preserved on the mud floors. Houses were large, 10–12 m (40–50 ft) long and 3–4 m (10–14 ft) wide. Low platforms of clay were built on the floors for the hearths. Different activity areas and trash dumps were apparent from the distribution of artifacts and other materials at the site (Figure 11.16). The food remains are remarkable; millions of fruit pits, nutshells, and seeds have been collected. The excellent preservation provides an unusual opportunity to reconstruct the complete diet of a prehistoric people. The diet of the inhabitants of Charavines was varied and nutritious. The occupants practiced farming, collecting, hunting, and fishing to obtain the foods they needed. A sizable portion of the diet came from wild sources. The animal bones at Charavines provide evidence of meat in the diet.
Chapter 11 Prehistoric Europe
The bones from the later phase of settlement represent at least 504 animals from numerous species (Table 11.2). Although there are several domesticated animals, they represent fewer than 10% of the animals in the diet by weight. Wild animals provided the bulk of the meat. The plant foods included a variety of dried or charred grains, seeds, nuts, berries, and fruits. More than 80 species of plants were found in the deposits at Charavines, including 17 species of weeds. Cultivated plants included three kinds of wheat, as well as barley, flax, and poppy. Several herbs, including oregano and thyme, were present. Fruits, nuts, and berries included wild apples, strawberries, plums, sloe berries, acorns, walnuts, hazelnuts, pine nuts, and beech seeds. Other wild plants included carrots, grapes, a material used as tinder for starting fires, and wild roses.
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Food-preparation techniques can be determined from the preserved remains. Whole dried apples are preserved in the deposits at the site, along with baked breads and other materials (Figure 11.17). Cooking utensils such as wooden spoons and whisks, and baking stones for making bread, were also found. Both gardens and fields were used for growing plants. Small gardens next to the houses were used for herbs, spices, medicinal plants, and some vegetables. Agricultural fields were used for cereal crops of wheat and barley. Cultivating tools such as antler hoes were found in large numbers at the site. A great deal of information on the technology of the inhabitants is preserved in the submerged layers at Charavines. The abundant wooden objects include bows and arrows; handles for axes, knives (Figure 11.18), and other tools; canoes and paddles; and planks for construction. Textiles and tools for cloth making are preserved as well, including spindle whorls for spinning yarn from wool and flax fiber, wooden combs used as weaving tools, and bone needles for sewing. Numerous pieces of rope and string were also found, as were beads and pendants of shell and various stones. Other items at Charavines document long-distance trade during the Neolithic. Some of the flint used for daggers came from a source known as
Page 523
Grand Pressigny in northern France. An amber bead must have originally come from the shores of the Baltic, hundreds of kilometers to the north. Stone for a polished axe was brought from Switzerland or southwestern Germany. Clearly, there were connections for trade and exchange across large areas of Europe, even during the Neolithic. Prehistoric sites with preservation such as that at Charavines are rare and for that reason are all the more important. Certainly, the information gained in the recovery of this underwater treasure was well worth the time, effort, and money expended.
Figure 11.17 Apples dried for storage. They were preserved among the many organic materials from Charavines in the wet lake sediments.
Figure 11.18 Two hafted flint knives. These extraordinarily rare examples show the craftsmanship involved in the technology of the Neolithic. The haft is wood, lashed to the flint blade with strands of bark fiber.
_ Site Charavines
523
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 524
Concept
<
The Megaliths of Western Europe Trademark tombs of the first farmers
Figure 11.19 The distribution of megalithic tombs (shaded areas) in western Europe and the location of Stonehenge, England, and Carnac, France. Notice that there are concentrations of tombs in some areas and a relative absence in other regions. Generally, the distribution of the tombs follows the Atlantic and North Sea coasts of Europe.
NORTH SEA
ATLANTIC
Stonehenge
OCEAN
Carnac
MEDITERRANEAN SEA 0
100
0 100
megalith A large stone monument.
_ _
menhir A large standing stone, found either alone or collectively in lines.
524
200 300 Miles 300 Kilometers
Agriculture spread to western and northern Europe at the beginning of the fourth millennium B.C. Almost immediately, farming societies began to erect structures made of massive stone slabs and boulders. These megaliths, as they are called, usually involved a burial area in or on the ground, surrounded by a chamber made of huge stones laid on top of one another without mortar. The entire stone tomb was then buried beneath a mound of earth to create an artificial cave. Often, a covered passage at the edge of the mound provided an entrance for later use of the tomb. Tens of thousands of these structures are found along the western fringe of Europe, still highly visible today, in Portugal, Spain, France, Belgium, Ireland, Britain, the Netherlands, Germany, Denmark, and Sweden (Figure 11.19). The megaliths are distributed geographically in a curious,
Chapter 11 Prehistoric Europe
patchy pattern that defies current explanation. Different traditions of pottery and house construction were associated with these monuments; apparently, different groups of people built the same kinds of tombs. The absence of metal objects in these structures makes it clear that they predate the Bronze Age in western Europe. Radiocarbon dates indicate an age between 4000 and 2000 B.C., during the Neolithic. Most of the megalithic structures were built early in this period. Varying greatly in size and in the number of stones used for construction, they were all built to withstand the test of time—to last for many, many generations (Figure 11.20). The megaliths fall into three major categories: menhirs, henges, and tombs. Menhirs are large standing stones, erected either singly or collectively in linear arrangements. Standing stones, either selected as long narrow blocks or
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
shaped intentionally to be set upright in the ground, occur in various heights, usually in the range of 1–5 m (3–15 ft). The largest known menhir comes from the town of Locmariaquer on the peninsula of Brittany in northwestern France. It now lies on the ground, broken into five huge pieces. Originally, this menhir was 23 m (75 ft) tall, the height of a six-story building, and weighed at least 350 tons. Perhaps the most impressive of the linear arrangements of these stones is found at Carnac, also in Brittany, where approximately 3000 large stones have been arranged in 13 parallel lines, stretching almost 6 km (4 mi) across the landscape (Figure 11.21). The stones are smaller at the eastern end, around 1 m (3 ft) high, and reach up to 4 m (13 ft) at the western end. The purpose of these linear arrangements is unclear, but they may have been intended to measure the cycles of the moon and predict eclipses. What is intriguing is that a smaller arrangement of wooden poles could easily have been used in place of such massive stone sentinels. Henge monuments, or circles, are defined by an enclosure, usually a cir-
Page 525
cular ditch and bank system, up to 500 m (1600 ft) in diameter. Not all henges contain stones; some appear to have been large timber structures. Stone circles, found primarily in the British Isles, are a special form of alignment with a definite astronomical significance. The best known of these is at Stonehenge (see p. 528), but hundreds of other stone circles dot the landscape of northern England and Scotland.
Figure 11.20 A megalithic tomb, or dolmen, in Denmark. The grave itself lies under the capstone. Many of these dolmens and other megalithic tombs would have been buried under a mound of earth.
Figure 11.21 The largest series of aligned menhirs in Europe at Carnac. There are approximately 3000 standing stones at the site in a series of rows. The standing stones are usually larger at one end of the row than at the other.
henge A monument defined by the
presence of an enclosure, usually made by a circular ditch and bank system, up to 500 m in diameter.
Concept The Megaliths of Western Europe
525
_
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
(a)
(b)
Figure 11.22 The three major types of megalithic tombs in western Europe: (a) dolmen, (b) gallery grave, and (c) passage grave. The chronological and geographic relationships among these tomb types are not clear.
dolmen A generic term for a megalithic tomb or chamber with a roof. passage grave A megalithic tomb
entered via a long, low, narrow passage that opens into a wider room, generally near the center of the structure.
_ _
gallery grave A megalithic tomb lacking an entrance passage; the burial room or rooms form the entire internal structure.
526
Page 526
(c)
A dolmen is a small megalithic tomb or chamber with a roof. Large stones and piles of earth were used to create these chambers. In spite of the enormous amount of labor required to obtain the stones and move them to the site of the tomb, the entire structure was often buried under a mound of earth. Megalithic tombs range from small, single-chamber structures to enormous hills of rock and soil that may hide a number of rooms and crawlways. Passage and gallery graves are two types of these larger tombs. A passage grave is entered via a long, low, narrow passage that opens into a wider room, generally near the center of the structure. A gallery grave, or
Chapter 11 Prehistoric Europe
long tomb, lacks an entrance passage, and the burial room or rooms form the entire internal structure (Figure 11.22). Some of the tombs had a movable stone door that could be opened to permit the interment of new bodies. Remains of the previous occupants were pushed to the sides to make room. These tombs may have been intended for all the members of a related group of farmsteads or hamlets; the tomb symbolized the collective and cooperative nature of the group. The earliest examples of megalithic tombs come from Brittany, dating to around 4000 B.C., the time of the transition to agriculture in this area. The first large tombs in Scandinavia at
10/1/09
12:13 PM
Page 527
© 2004 by Sidney Harris. Reprinted with permission of ScienceCartoonsPlus.com.
pri31057_ch11.pp506-559.qxd
about the same time were earthen long barrows, or mounds, with a single log tomb inside, apparently imitating the plan of a Neolithic house. These earthen structures, intended for the burial of a single individual, could be transformed into larger stone chambers for the collective burial of several individuals. These tombs served an important purpose for the living as well. Such “cults of the dead” likely involved both ancestor worship and property rights. Elaborate burial rituals and monument construction integrated the cult. Ancestor veneration may have supported claims to agricultural fields for local communities of farmers. The construction of a permanent burial monument provides dramatic evidence of one’s tenure in place and the inheritance of rights to the land. Part of the ritual associated with a cult of the dead apparently dictated that burials be placed in large stone tombs after lengthy ceremonies and activities elsewhere. Only the bones of the deceased were placed in the tombs; the flesh had decomposed or had been removed before entombment. The tombs do not appear to have been major repositories for wealth or elaborate furnishings for the dead. Grave goods are rare; usually, only a few pots, stone tools, and animal bones are
found. There is, however, evidence that pottery vessels with food and drink were regularly placed at the entrance of the tombs in ceremonial offerings for the deceased. Megalithic structures provide dramatic and enduring evidence of the impact of agriculture on the inhabitants of western Europe. Construction of monumental architecture is one clue to the increasing complexity of societies in this area. Shortly after farming was adopted, a pronounced trend toward regionalization began. In spite of the evidence of a widespread cult such as megalithic burials, various regional styles in material culture arose in conjunction with the fortification of settlements and increasing evidence for warfare. It appears that more distinct and delineated societies were identifying themselves across the continent, and this pattern continued into the succeeding Bronze Age and Iron Age.
Megalithic tombs represent a statement placed on the landscape by Neolithic peoples. What is that statement?
y
barrow An earthen mound covering
a burial.
Concept The Megaliths of Western Europe
_
527
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 528
Site
:
Stonehenge A temple to the sun on England’s Salisbury Plain The most impressive prehistoric monument in the British Isles, and perhaps all of Europe, is Stonehenge on the Salisbury Plain in southern England. It is not the largest or the richest, but Stonehenge holds an enormous fascination, as much for the mysterious aura that surrounds it as for the impressive feat of construction. Some of the stones have been taken away over the centuries; many have fallen and lie half-buried in the earth. Other stark, brooding, gray stones still stand, arranged in the circles and arches that outline this “observatory” and ceremonial center of late Neolithic and Bronze Age Britain (Figure 11.23). The construction and elaboration of Stonehenge took place over a period of 1500 years, from approximately 3000
Figure 11.23 An aerial view of Stonehenge, with the Aubrey holes in the background. Note also the circular ditch around the outside of the monument.
_ _
trilithon A massive stone lintel.
528
Chapter 11 Prehistoric Europe
to 1500 B.C. On the one hand, Stonehenge was rebuilt almost continuously during this period, changing in purpose and function. On the other hand, the previous constructions remained visible to subsequent generations and were respected as part of the continuity of the monument from beginning to end. This area of the Salisbury Plain must have held as great a fascination for its prehistoric inhabitants as it holds for modern tourists. In addition to the focus on Stonehenge, the landscape around the site is filled with tombs and monuments to the dead. The earliest stage of the monument was built between 3000 and 2900 B.C. It consisted of a circular bank and ditch, almost 100 m (330 ft) in diameter, cut into the chalk of the subsoil.
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 529
The bank on the inside of the circle would have stood almost 2 m (6 ft) high, piled up with material dug from the ditch. It is probable that a wooden palisade was erected along the bank in the fashion of other causewayed enclosures known in England at this time. A broad entrance to the circle lay on the northeast side, and two smaller ones lay to the south. Inside this circle and bank were dug the 56 “Aubrey” holes, regularly spaced around the perimeter, named after their seventeenthcentury discoverer. These holes were filled shortly after their original excavation, and their original purpose is unknown. During a second stage of construction, until 2500 B.C., the monument was extensively remodeled. The major addition was a circle of upright timbers or posts in the center of the monument. Similar structures, known as wood henges, from this period have been found in other parts of the Salisbury Plain. Two entrances were elaborated, and one of the southern entrances was closed. A short “avenue” of wooden posts was added toward the southern entrance, and a substantial timber screen was built across it to block the view to the center. Cremation burials were placed in a few of the depressions and also in the bank and ditch areas around the perimeter of the monument. The third stage of construction involved a series of major additions between 2500 and 1600 B.C. The timber
circle at the center of the monument was replaced with stone in a series of episodes. A circle of carefully shaped pillars and lintels of Sarsen sandstone was erected in the center (Figure 11.24). Huge columns of this sandstone were quarried as far away as 30 km (20 mi), shaped into pillars, and dragged to the site. Five trilithons (two upright pillars with a crosspiece on top) of Sarsen stone were erected in a horseshoe-shaped arrangement open toward the northeastern entrance. One of these trilithons is distinctly taller than the other four, standing 8 m (25 ft) high. The lintel on this trilithon is 5 m (16 ft) long and over 1 m (3 ft) thick. The larger Sarsen pillars weigh as much as 50 tons (equivalent to a loaded train car) and were likely moved on oak rollers. A system of scaffolding was probably used to raise the stones into position. The individuals involved in the various construction phases of Stonehenge must have numbered in the hundreds if not thousands. Two parallel, crescent-shaped rows of standing bluestones were erected in the center of the circle, along with several upright stones. This blue-tinted volcanic dolerite came from a source some 250 km (150 mi) distant in the Prescelly Mountains of southern Wales (Figure 11.25). Certainly, part of their journey was by water, but the movement of such large stones was a major accomplishment in itself. More standing stones were added near the entrance and inside the periphery of the ditch. Slightly later, a new circle of
Probable route for transport of Prescelly Milford Mts. Haven
bluestones sarsens Marlborough Downs
W A L E S Av
Bristol 0
on R ive r
Avebury Devizes
Chann
el
25 Miles
Stonehenge 0
25 Kilometers
Figure 11.24 Construction technique at Stonehenge. Wooden scaffolding was probably used to raise the lintel stone into place. The amount of labor and materials involved in the construction process reflects a major undertaking. Figure 11.25 Probable routes of transport for bluestones from the Prescelly Mountains in Wales and Sarsen stone from near Avebury to Stonehenge. Site Stonehenge
529
_
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 530
Stonehenge A
Stonehenge B Entrance
Zone of postholes
Screen “Avenue”
Entrance Entrance Stonehenge C
Stonehenge D Avenue
N
0
Figure 11.26 Four major stages in the construction of Stonehenge: (a) a causewayed enclosure with an internal palisade, c. 3000–2900 B.C.; (b) a timber construction with avenue and screen and central zone of postholes, 2900–2500 B.C.; (c) the major stone structure, 2500–1600 B.C.; and (d) the final phase with two rings of pits surrounding the central stones, 1600–1500 B.C.
_ _
530
standing bluestones was placed inside the horseshoe of trilithons. In the final stage of construction, a ring of bluestone pillars was raised inside the Sarsen circle, but outside the horseshoe. In addition, two rings of large holes were dug around the outer circle. These may have held standing stones that are now missing (Figure 11.26). An “avenue” some 500 m (1600 ft) long and 15 m (50 ft) wide, flanked on either side by an earthen ditch and bank, was constructed running to the northeast from the circle of Stonehenge, leading to the River Avon. This
Chapter 11 Prehistoric Europe
30m
avenue runs through one of the densest concentrations of prehistoric burial monuments in England. The Heel Stone was brought to the site at this time and erected in the avenue about 10 m (33 ft) from the northeastern edge of the circle. This irregular boulder rises almost 5 m (16 ft) above ground level and weighs at least 35 tons (equivalent to an 18wheel truck). Stonehenge functioned in part as an observatory to record the summer solstice. On the 21st or 22nd of June each year, the dawn sun rises directly over the Heel Stone. When the weather
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
is clear, the first sunlight passes across the Heel Stone and through the double standing stones at the entrance, bisects the two horseshoes of standing stones, and reaches the altar stone in the very center of the circle. Today, the sun just misses the exact top of the Heel Stone as a result of small changes in the earth’s axis since the monument’s construction. This shift was used in 1905 by the Royal Astronomer, Sir Norman Lockyer, to estimate the date of the monument as 1900 B.C., an accurate prediction made many years before other methods of absolute dating were available. Some have argued that Stonehenge was an astronomical computer, used to record various lunar and stellar alignments. One of the more popular theories is that the circle of stone and holes could have been used to predict lunar eclipses. There is no strong evidence, however, for any alignment other than that of the summer solstice. The last stage of Stonehenge reflects its declining use in the period between 1600 and 1500 B.C. Two concentric rings of pits were added around the stone circles in the middle of the monument. Their function is unknown. In addition, several carvings were added to the trilithons, including a depiction of a bronze dagger. New fieldwork has begun at Stonehenge in recent years, and the discoveries increase rapidly. Stonehenge was first built about the same time as the timber rings at Durrington Walls and Woodhenge. These wood henges are
Page 531
connected to Stonehenge by an avenue of parallel banks and ditches over a distance of 8 km (5 mi). In addition, the remains of a potentially very large Neolithic village may underlie the construction of the henges at Durrington Walls. Stonehenge served as an important burial ground from its earliest beginnings and for several hundred years. The scientists in charge of the new excavations, Mike Parker Pearson and his colleagues, believe that this avenue connected the living (wood) with the dead (stone) domains in this sacred landscape. Preserving Stonehenge has been a problem for British archaeology in recent years. Although it is a mecca for tourists to England, just behind the Tower of London in popularity, the wear and tear wrought by visitors has led to more careful regulation of the monument grounds. Years of postwar tourism had dire consequences— inquisitive fingers wore down the engravings on the stones, and the weight borne by millions of feet eroded the ground down to the chalk bedrock. At the summer solstice in late June, huge festivals celebrated the summer as present-day Druids welcomed the sunrise. Today, most of the monument is not directly accessible; walkways direct the visitor past the mute sentinels of the past. While such means of protection distance the visitor from the stones, they do serve to help ensure that Stonehenge and its heritage will remain for the future as a monument to and from the ancient Britons.
What is Stonehenge: it is the roofless past; Man’s ruinous myth; his uninterred adoring Of the unknown in sunrise cold and red; His quest of stars that arch his doomed exploring. And what is Time but shadows that were cast By these storm-sculptured stones while centuries fled? The stones remain; their stillness can outlast The skies of history hurrying overhead. —Siegfried Sassoon (1961)
_ Site Stonehenge
531
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 532
Concept
<
The Aegean Bronze Age Home of the heroes of Homer’s legends
GREECE
Troy
AEGEAN SEA TURKEY Thebes Corinth Mycenae
Athens Tiryns
Pylos
Franchthi Cave
MELOS MEDITERRANEAN SEA Knossos
Mallia Phaestos
0 0
100 100
200 Kilometers
Figure 11.27 Locations of important Minoan and Mycenaean sites in the Aegean region. Only a small number of the Aegean islands are shown, including Melos, the source of obsidian during the Mesolithic.
_ _
532
CRETE
200 Miles
The wine-dark waters of the Aegean Sea bathe the shores of Turkey to the east and Greece to the north and west. The long, mountainous island of Crete marks the southern limits of this sea. The Aegean is splattered with small, rocky, volcanic islands, actually the summits of a submerged mountain range, with thin, poor soils and a dry climate (Figure 11.27). These rugged, barren conditions may have been a hidden benefit for the early inhabitants. The absence of large areas of fertile farmland meant that
Chapter 11 Prehistoric Europe
crops other than cereals had to be cultivated. Wheat could be grown in more sheltered areas with deeper soils, but grapes, olive trees, and sheep flourished on the rocky slopes of the islands. Olives are remarkably nutritious and provide oil that can be burned, eaten, cooked, or rubbed on the skin and hair. Not only are grapes a delicious fruit, but their fermented juice is wine. Oil, wine, and wool became important exports for the economies of these small islands. The inhabitants must have also relied on the sea for fish and other foods; sailing and a knowledge of the sea would have been an essential part of life. Seafaring and trade permitted the movement of goods and foods between islands, and between the islands and the mainland. The strategic location of Crete and the Aegean islands along the main avenues of sea trade enabled the inhabitants to become the middlemen in moving goods between the civilizations of Egypt and Southwest Asia and the settlements of Europe. The demand for wine, olive oil, pottery, textiles, and other goods enhanced the economic well-being of the people of the Aegean islands. A pattern was apparently established rather early in which raw materials were taken to the islands and made into finished products. Craftworkers used the potter’s wheel to make fine ceramic vessels. Others carved stone, bone, and ivory seals for marking economic transactions; produced wooden tools; sculpted figurines and bowls of marble, obsidian, and other colorful stones; or created jewelry and other luxury items. After 3000 B.C., the craftworkers of the Aegean also began to make objects of metal: bronze, silver, and gold. Smiths produced bronze tools and weapons by the thousands for both
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 533
(b)
(a)
(c)
Figure 11.28 Pictographic scripts, used in early writing in the Aegean area. These tablets are from Crete and show both Linear A and Linear B systems: (a) Linear A, with a four-wheeled wagon; (b) Linear B, with sword, chariot, and horse; (c) Linear B, describing the issue of a set of armor; and (d) Linear B, counting gold cups and bull’s-head drinking vessels. These texts often recorded the goods and equipment owed by the local populace to the palace, early tax records. (d)
local use and export. This metal was the signature material of the Bronze Age, and its production was one of the primary reasons for the explosion of economic and political power in the Aegean region. Bronze was discovered, probably by accident, shortly before 3000 B.C. Bronze is a mixture of copper and tin or arsenic. Copper ores often contain some arsenic naturally. In all likelihood, early smiths discovered that those ores that included arsenic produced a slightly harder material that was easier to cast. Further experimentation must have led to the discovery of tin as another alloy, and bronze metallurgy had begun. Bronze has several advantages over copper. It can be recycled repeatedly, whereas copper loses its tensile strength in recasting. Bronze also holds an edge much better than
(Pictographic scripts from Greece in the Bronze Age by Emily Vermeule. Copyright © University of Chicago Press. Used with permission of the publisher, University of Chicago Press.)
copper, and most of the early bronze objects were weapons: swords, daggers, spearheads, and arrowheads. The advent of metallurgy, first of copper and then of bronze, silver, and gold, greatly increased trade and the movement of goods throughout the Aegean region and the rest of Europe. Much of Europe was involved in the movement of materials and goods into the economic magnet of the eastern Mediterranean. The Aegean Bronze Age dates to about 3000–1000 B.C., ending with the beginning of the Iron Age and the rise of the classical Greek civilization of Plato and Homer. There were two major centers of development and power in the Aegean: one on Crete and one on mainland Greece. The civilization that emerged on the island of Crete was known as the Minoan and reached
bronze A hard, durable metal made from a mixture of copper and tin or arsenic.
Concept The Aegean Bronze Age
533
_
pri31057_ch11.pp506-559.qxd
10/1/09
citadel A hilltop fortress, the charac-
_ _
teristic settlement of the ruling elite of Mycenaean civilization, 1700–1100 B.C.
534
12:13 PM
Page 534
its peak between 2000 and 1450 B.C. During this period, the Minoans dominated the Aegean through sea power and the control of trade in the eastern Mediterranean. The seats of power on Crete were palaces and villas, residences of the local rulers who directed this early state. Defensive fortifications were not needed by the islanders; they were protected by their ships. Three writing systems were used on Crete, including an early hieroglyphic system, replaced around 1700 B . C . by a system known as Linear A, which is still undeciphered. A second writing system, Linear B, has been decoded and is related to the archaic Greek language (Figure 11.28). Linear B was developed on mainland Greece and later introduced to Crete by the Mycenaeans. The Mycenaeans on mainland Greece controlled most of the Aegean between 1600 and 1100 B.C. and took over Crete after 1450 B.C. The Mycenaean civilization was dominated by a series of hilltop fortresses, or citadels, interconnected by roads. These citadels were ruled by powerful warrior-kings, whose graves are among the richest ever uncovered in Europe. Episodic alliances among the citadels led to greater political, economic, and military power, and the Mycenaeans became the major force in the Aegean around 1500 B.C. The collapse of Mycenaean power and the abandonment of the heavily fortified citadels after 1100 B.C. is one of the more intriguing mysteries of Aegean archaeology. The Bronze Age in the Aegean marks a watershed in the prehistory of Europe. In the Mediterranean Basin, early civilizations rose in the Aegean, on mainland Greece, and later on the peninsula of Italy. These societies were literate and ruled by kings; they inhabited large towns, maintained armies and navies, collected taxes, and established laws. These states controlled trade over large areas and extracted a variety of raw materials and other products from the rest of Europe. North of the Alps, there was less political integration; societies operated at a tribal or a chiefdom level, on
Chapter 11 Prehistoric Europe
a smaller scale. This pattern continued essentially until the Roman conquest of France and much of Britain, shortly before the birth of Christ. The Romans introduced writing systems and true statecraft into northwestern Europe for the first time. Prehistory ended much earlier south of the Alps, where the literate civilizations of Greece, in the first half of the first millennium B.C., and of Rome, in the second half, dominated the Mediterranean Basin.
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 535
Site
:
Knossos The mythical halls of the Minotaur on the island of Crete
Sir Arthur Evans, then keeper of the Ashmolean Museum in Oxford, England, traveled to Crete in 1894 and “discovered” an extensive group of ruins, buried under a low mound of soil and collapsed walls, at a place known as Knossos (kuh-NOS-sus). Of course, the local people had known about these ruins for millennia, but his was the first report back to the Englishspeaking world. Beginning in 1900, Evans spent the remaining 35 years of his life excavating at Knossos. He restored many of the areas he had ex-
cavated, rebuilding the walls and repainting the plaster in the vivid colors that had been preserved in the ruins. Today, the palace that he uncovered is a monument to his labor and his vision of the restoration (Figure 11.29). The first Bronze Age palace at Knossos was erected around 3000 B.C. on top of 7 m (22 ft) of Neolithic deposits that had accumulated for several thousand years. A series of palaces were built one on top of another, each larger and more elaborate, as the settlement and its administrative structure grew.
w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the Minoan ruins at Knossos, see the Internet exercises on your online learning center.
Figure 11.29 A portion of the reconstructed palace at Knossos, Crete, showing a street and the facade of a building. Large parts of the central area of the site were extensively reconstructed by the excavator Arthur Evans.
_ Site Knossos
535
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Figure 11.30 The plan of Knossos around 1900 B.C., including the palace, several adjacent mansions, smaller houses, and other buildings, along with a system of paved roads. The original settlement at this location was a Neolithic village at the junction of the two streams.
Page 536
Deposits
Little Palace
Royal Villa
Armory West Royal Road
North House
Royal Road
North Pillar Hall
Grandstand Northeast House
House of Frescoes
East House Southeast House
House Stepped of the portico Chancel
rse cou d l O tream of s Vlychia
THE PALACE
Cen t courral t
of rse d cou Roa Old noan Mi
West court
Houses
Stream Viaduct
Hill of Gypsades
Houses
Gypsum quarry
Kairatos Stream
Caravanserai
Garden 0
100 m
Houses
_ _
pithos A large clay storage jar.
536
Knossos covers almost 25,000 sq m (6 acres, an area the size of a large basketball arena), in a complex of buildings and construction that included the palace itself, several surrounding large mansions, and many smaller houses, connected by roads (Figure 11.30). The palaces were the centers of the Minoan state, laid out and built according to plan, with a large, rectangular central court surrounded by myriad rooms, numerous corridors, a maze of courtyards, grand staircases, private apartments, administrative chambers, enormous storerooms, baths, and even a sophisticated plumbing system (Figure 11.31). The complex of rooms and buildings housed many of the administrative, economic, and religious func-
Chapter 11 Prehistoric Europe
tions of the government. Many important craft workshops were also located in the palace or housed in adjacent buildings. Functional space was carefully designed and separated according to residential, administrative, storage, religious, and manufacturing uses. The palace was a multistory building with extensive storerooms. Long, narrow rooms held enormous storage jars, or pithoi, along the walls for oils, wine, and other liquids, and stonelined pits in the floor were filled with wheat and other cereals (Figure 11.32). It is clear from this arrangement that the palace complex controlled the economic activities of the state. Frescoes and murals decorated the walls of the palace, depicting various
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 537
Figure 11.31 An artist’s reconstruction of the palace at Knossos, showing the multistory building, the large central plaza, and the open nature of the construction.
aspects of Minoan life. Shrines are scattered throughout the palace as well, documenting the integration of church and state in the early Aegean civilizations. Much attention has been focused on the open-bodice costumes of the goddesses and the acrobatic bullfighters who are often depicted. Religious ceremonies appear to have combined several elements, including the bull, a sacred axe, and snakes. The extensive network of trade that was directed from Knossos and other Minoan centers is evidenced by the variety of raw materials found in the palace: copper from Cyprus and Turkey; ivory, amethyst, carnelian, and gold from Egypt; lapis lazuli from Afghanistan; and amber from Scandinavia. These commodities, and the Cretan ships that carried them, were the foundation of the wealth and power of the Minoan state. The Egyptians feared the “great green sea,” but the sturdy Minoan ships, with their deep keels and high prows, weathered the storms of the Mediterranean and controlled the sea lanes. The palace of Knossos was destroyed at least twice during its history. The first destruction, in 1700 B.C., was marked by extensive wall collapse and evidence of a major fire. Many of
the other palaces and villas on Crete show evidence of similar destruction at the same time; it seems clear that a major earthquake must have occurred. Most of the palaces were rebuilt following this episode of destruction. The second period of destruction at Knossos dates to approximately 1450 B.C. and marks the end of the Minoan civilization. The palace was reoccupied following this episode, but the pottery and other artifacts indicate that Mycenaeans from the mainland were in control. Other palaces on Crete also were destroyed at this time, but not simultaneously, as in 1750 B.C. The reason for this last episode of destruction and the collapse of the Minoans is less clear but probably involves the growing power of the Mycenaeans.
Figure 11.32 Basement storerooms at Knossos. These palace basements were warehouses with storage pits in the floor for grains and huge storage jars, or pithoi, for liquids along the walls. Such storage areas reflect the economic primacy of the palace. Site Knossos
537
_
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 538
Site
:
Mycenae Fortress of the warrior-kings of Bronze Age Greece The Bronze Age in Europe, and its accompanying weaponry, ushered in a period of conflict and warfare in which the skills, attitude, and power of the military seemed to take precedence over other aspects of society. A new warrior class emerged during this period; weapons and armor were the primary burial goods, and martial and hunting scenes dominated decorative art. A military presence is strongly visible in the Bronze Age citadels of southern Greece, the “halls of the heroes.” These are the people, so vividly described in The Iliad of Homer, who sailed to Troy, on the west coast of Turkey, and eventually sacked the city around 1250 B.C. This is the civilization of Agamemnon and Ulysses. These early Greeks gradually wrested power away from Crete and the Minoans and
Figure 11.33 An aerial photograph of the citadel of Mycenae in Greece. The fortresses of the Mycenaeans usually were situated in highly defensible locations.
_ _
shaft grave A vertical tunnel cut into rock and holding the tombs of Mycenaean elite.
538
Chapter 11 Prehistoric Europe
came to dominate the Aegean between 1600 and 1100 B.C., a time known as the Mycenaean period in prehistoric archaeology. The citadels, of which the site of Mycenae (my-SEEN-ee) is best known, were fortified palace towns, located on high, defensible points on the landscape (Figure 11.33). The early rulers of the citadel of Mycenae were buried in shaft graves, straight-sided pits 6–8 m (20–25 ft) deep, cut into the soft rock of their hilltop settlement. Groups of shaft graves were enclosed in a circle of standing limestone slabs (Figure 11.34). Two such grave circles have been excavated at Mycenae, and several tombs were found in each shaft. The walls of the tombs were lined with brick or stone, and the entire structure was covered with a timber roof. At a
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
later date, the shaft was reopened and another tomb added. Nineteen individuals were buried in the six shafts of grave circle A at Mycenae, two to five people in each shaft. There were nine men, eight women, and two children. The grave goods from Mycenae are among the most spectacular finds from the Bronze Age. The graves included precious metals and stone in the form of weapons, vessels, masks, and other objects (Figure 11.35). One early grave in these tombs contained more than 5 kg (11 lb) of gold. There are gold and bronze masks and drinking cups, necklaces, earrings, a crystal bowl carved in the shape of a goose, and swords and daggers with gold and lapis inlays. Ninety swords were found in the graves of three individuals. Amber from northern Europe, ivory from Africa, silver from Crete, glass from Egypt, and great amounts of gold were entombed along with these early rulers. In addition, the corpses were apparently covered with hundreds of leaves, flowers, butterflies, and stars cut from thin sheets of gold.
Page 539
Heinrich Schliemann, excavator of the graves, described opening one of the shafts in a letter to a friend in 1876: There are in all five tombs, in the smallest of which I found yesterday the bones of a man and a woman covered by at least five kilograms of jewels of pure gold, with the most wonderful archaic, impressed ornaments; even the smallest leaf is covered with them. To make only a superficial description of the treasure would require more than a week. Today I emptied the tomb and still gathered more than 6/10 kilogram of beautifully ornamented gold leafs; also many earrings and ornaments representing an altar with two birds. . . . There were also found two scepters with wonderfully chiselled crystal handles and many large bronze vessels and many gold vessels. (Quoted in Vermeule, 1972, p. 86)
A kind of radiance, like that of the sun or moon, lit up the high roofed halls of the great king . . . the interior of the well-built mansion was guarded by golden doors hung on posts of silver which sprang from the bronze threshold. . . . On the other side stood gold and silver dogs . . . to keep watch over the palace. —Homer, describing the home of a legendary Greek king
Schliemann was convinced that he had found the gold death mask and body of Agamemnon, who was murdered by his wife’s lover when he returned from the conquest of Troy. Although that individual was probably not
Figure 11.34 The massive outer walls of the citadel, enclosing one of the shaft-grave circles and several of the elite burials at the site. The grave shafts were originally filled and covered in the construction of the burial ground.
_ Site Mycenae
539
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 540
Figure 11.35 A gold suit placed over a buried child at Mycenae. Gold masks were placed on the faces of several of the individuals buried here and at other Mycenaean sites.
Not even those who lived long ago before us and were sons of our lords, the gods, themselves half-divine, came to an old age and the end of their days without hardship and danger, nor did they live forever. —Simonides, in Vermeule (1972)
tholos A large, beehive-shaped tomb,
constructed using the corbel arch technique, characteristic of the Mycenaean civilization of Greece. Cyclopean A term describing the huge
_ _
stone walls of Mycenaean tombs and fortresses; from Cyclops, the mythical giant.
540
Agamemnon, and the mask may well have been a forgery brought in by Schliemann, the contents of the graves of Mycenae clearly document the wealth of this civilization and its rulers. By 1400 B.C., a new kind of grave, the huge, vaulted, beehive-shaped tholos tomb, was constructed for the major rulers. Several of these architectural wonders still stand today. The Treasury of Atreus at Mycenae is the finest example of a tholos tomb. The roof of the vault stands more than 13 m (40 ft) above the floor, which is 15 m (50 ft) in diameter. The dramatic doorway to the tomb is 5 m (16 ft) high, and the lintel across the door weighs more than 100 tons (Figure 11.36). Unfortunately, the contents of the tomb were stolen long ago. Many large and small settlements from this time period were scattered across southern Greece. A sophisticated system of graveled roads for chariots and carts, with stone bridges and culverts, connected the towns and
Chapter 11 Prehistoric Europe
villages. Mycenaean towns were heavily fortified with stone walls known as Cyclopean walls because of their massive size. Major citadels from this period are known from Mycenae itself: Tiryns, Pylos, and Thebes. At the citadel of Tiryns, the great walls are 15 m (50 ft) thick and contain internal passages. At Mycenae, the hilltop was leveled and terraced to accommodate the walls of the fortress as well as the inhabitants of the palace and town. A long, narrow road, flanked by high stone walls, leads up to the Lion Gate, the entrance to the citadel, so named because of the enormous stone sculpture of two lions that crowns the gate. The monumental stone walls of Mycenae encircle an area 1100 m (3500 ft) in diameter (the equivalent of two city blocks), enclosing the palace of the king as well as a number of residences and other structures. The Mycenaean palaces combined many of the administrative, military, and manufacturing functions of the kingdom within the residence of the ruler. Workshops for crafts, guardrooms, storerooms, and kitchens were attached to the rear of the palace. There was a small postern gate at the back of the citadel. During a siege, fresh water was available from a cistern, located at the bottom of a rockcut tunnel and staircase deep inside the hill of Mycenae. The surrounding villages supplied plant foods and meat, men, and materials to the lord of the citadel. This information comes from preserved clay tablets with Linear B script. In some instances, these soft clay tablets were accidentally burned in fires that swept the citadels and hardened the clay, thereby preserving the script. The subjects of the texts are primarily economic, dealing with inventories, shipments, and quotas of items to be paid to the palace in tribute. At the palace of Pylos, where a major hoard of tablets was preserved, a list of different occupations in the kingdom was recorded: bakers, bronzesmiths, carpenters, heralds, masons,
pri31057_ch11.pp506-559.qxd
10/1/09
12:13 PM
Page 541
Figure 11.36 The entrance to a tholos tomb, in which the later kings of Mycenae were buried. This example at Mycenae is called the Treasury of Atreus and dates to 1400 B.C. These monumental constructions were architectural wonders both then and now.
messengers, potters, shepherds, and “unguent boilers.” Many other skills are also evidenced in the artifacts and architecture of the citadel, such as delicate ivory carving, fresco painting, metal inlaying, and arms manufacturing. The reason for the collapse of Mycenaean civilization remains a mystery. Unrelated to drought or outside conquest, perhaps it was simply the result of a culmination of centuries of conflict and competition. After 1100 B.C., Athens began to assert its importance in Greece as the center of the new industry of ironworking, and the citadels of the Mycenaeans fell into disuse. Power shifted for a brief period
to the north, to Macedonia, the realm of Philip and his son Alexander the Great. The Iron Age civilization of classical Greece blossomed from 600 to 300 B.C. and gave rise to the golden age of Athens, the exploits of Pericles, the wisdom of Socrates and Plato, the inventions of Archimedes, the writings of Homer, and the foundations of much of Western civilization.
_ Site Mycenae
541
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 542
Concept
<
The Bronze Age North of the Alps Innovation and growing populations One of the more pronounced trends in the European Neolithic was regionalization, the development of distinctly local traditions. Initial farming cultures expanded over broad regions (Figure 11.37). Settlements were generally located in open and unprotected spaces, and pottery styles were similar across very large areas. Very quickly, however, population growth and the development of permanent field systems resulted in competition and conflict between groups. By 3000 B.C., the continent was occupied by well-entrenched farming populations making stone tools and pottery—cultivating, trading, and fighting. Later Neolithic settlements were often located in defensible positions and heavily fortified. Pottery styles became more limited in their distribution. At the same time, trade and exchange expanded in scope. A variety of materials and finished goods were moved long distances across Europe. Obtaining raw materials, manufacturing trade items, and transporting finished goods became an important part of Neolithic economic systems. Flint, for example, was mined in Denmark, Belgium, England, and elsewhere and polished into fine axes for trade.
Figure 11.37 A wooden plow from the Bronze Age of southern Scandinavia.
_ _
542
Chapter 11 Prehistoric Europe
Metals appeared later in the northern part of Europe than in the Aegean area. Copper first appeared north of the Alps around 4000 B.C., coming primarily from southeastern Europe. Bronze objects began to appear in graves and cemeteries of farming settlements in this region after 2000 B.C. By this time, Bronze Age cultures in the eastern Mediterranean had developed into powerful political entities through a combination of alliances, sea power, craft production, and the control of trade. The eastern Mediterranean acted like a magnet for the valued raw materials from the rest of Europe (Figure 11.38). Copper, tin, and gold from sources in Ireland and England moved across the English Channel, down the Seine and Rhone rivers in France, to the Mediterranean for shipment to the Aegean. Copper ores and ingots from the Carpathian Mountains in eastern Europe were brought overland through the Brenner Pass in the Alps to Greece. Amber from the Baltic and North Sea coasts of Denmark and Poland was imported to the Aegean. Other exports, such as furs and slaves, may have also been sent in exchange for finished bronze weapons, pottery vessels, and bronze and gold jewelry.
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 543
Figure 11.38 The major sources of gold, copper, tin, and amber in Europe and some of the trade routes to the Mediterranean. Three of the major centers of the Bronze Age north of the Alps were in southern England, Denmark, and the Czech Republic. The exchange of rare and valued materials between the Aegean and the rest of Europe gave rise to a number of powerful societies.
BALTIC NORTH
SEA
Danish Bronze Age
SEA
Wessex Unetice
Brenner Pass
MEDIT 0
1000 km
New things and ideas filtered back into Europe along these corridors of trade. The third millennium B . C . in northern and western Europe witnessed a number of major innovations, including bronze and gold, the oxcart and oxen as draft animals, the wheel, the horse and chariot, and new weapons. Secondary animal products such as milk and wool became more important and fostered new industries. Wealthy and powerful hierarchical societies arose at the centers of these important trade routes in the early Bronze Age after 2500 B.C. The Wessex culture in England elaborated the construction of Stonehenge and erected hundreds of burial mounds, or barrows, across southern England. The Unetice culture in the Czech Republic, known from several “princely” burials, dominated central Europe. The Bronze Age in Denmark and southern Sweden was spectacular in terms of the quantity of fine metal objects that were buried in many funerary mounds
ER
RA
NE
A
Mycenae N
SE
A
Crete
and caches. In fact, much of what is known about the Bronze Age comes from large barrows or caches of metal objects hidden in the ground; relatively few houses or settlements have been discovered or excavated. Such information provides a limited but spectacular view of only a small, wealthy segment of Bronze Age society north of the Alps.
_ Concept The Bronze Age North of the Alps
543
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 544
Site
:
Borum Eshøj A Bronze Age tomb in Denmark
Figure 11.39 Two Bronze Age barrows (burial mounds) along the coast of Denmark. These burial mounds dot the landscape of northern Europe and were constructed along roads and other routes of transport.
Figure 11.40 An example of a Bronze Age log coffin burial from Egtved, Denmark. In the hollow were placed blankets, a cowhide, two birchbark containers, tools, and a fully clothed young woman, wearing a string skirt, shirt, belt, and huge bronze belt buckle. The vagaries of preservation removed most traces of the body except for her hair.
_ _
544
Chapter 11 Prehistoric Europe
Bronze Age barrows dot the landscape of southern Scandinavia (Figure 11.39). In a few instances, the contents of these barrows have survived, providing a glimpse of the elite of Bronze Age society in northwestern Europe. Dressed in their finest clothing and jewelry, deceased individuals were placed in a coffin, a huge log of oak, split in half and hollowed inside (Figure 11.40). This coffin was then covered with a pile of stones and buried under a mound of cut sods and soil. The coverings of these tombs sometimes sealed the contents from the air, and the log coffins quickly filled with groundwater. Recent studies suggest that water may have been intentionally added during construction. These conditions preserved both the coffins and the contents to a remarkable extent. Borum Eshøj (BORE-um ES-hoy) was one of the largest Bronze Age barrows in Denmark, located slightly north
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 545
II I A
III
7
9
B 10
III
C
8
Figure 11.41 The ground plan of an excavated Bronze Age burial mound at Borum Eshøj in Denmark. The cross section at the top shows the stages of mound building and the location of the log coffins. The plan view shows the stone rings in the mound, the coffins, and other burial features.
III
I
13
A2
III 5
A1 1
7
A
8 9 B10
A3
11
4
15
12
C
16 14
2
Figure 11.42 A woman’s woolen tunic and belted skirt from one of the Bronze Age log coffins, one of the earliest examples of preserved clothing anywhere in the world.
3
III
II
N
II
0
10
of the modern city of Aarhus (Figure 11.41). The original mound was almost 9 m (30 ft) high (equivalent to a threestory building) and 40 m (130 ft) in diameter. The barrow was first opened in 1875 by the landowner, who was removing the rich soil of the mound to add to his fields. Three oak coffins were found in the mound, containing an elderly man, an elderly woman, and a younger man. The central log coffin was 3 m (10 ft) long and contained the body of a 50- to 60-year-old man, lying on a cowhide. He wore a wool cap over hair that was once blond, now stained black by tannic acid. His chin, also stained dark, was clean-shaven, and his teeth were in good condition. He was
20
30 m
dressed in a wool skirt with a rope belt. The body of the younger man, age 20, was found in another coffin. Wearing a wool shirt held together by a leather belt and a wooden button, he was buried with a bronze dagger in a wooden sword scabbard, a bone comb, a round bark container, and a wooden pin to fasten his cloak. The elderly woman was extremely well preserved. She was 50–60 years old and 1.57 m (5 ft, 2 in) tall. The first item found when the coffin was opened was a cowhide with the hairs still intact. Beneath the hide was a wool rug on top of the woman’s body. She was buried wearing a skirt and tunic of brown wool, and a tasseled belt (Figure
_ Site Borum Eshøj
545
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 546
Figure 11.43 The geographic distribution of Bronze Age barrows in Denmark. Note the distinct concentrations in certain areas and especially the linear patterns that appear in the southern half of Jutland, probably following old road systems.
11.42), as well as a hairnet of wool thread with her long hair still inside. A comb made of horn was found next to her hair. Various bronze objects in the coffin included a pin, a dagger with a horn handle, a belt disk, and rings for the fingers, arms, and neck. Bronze Age barrows were built for the wealthier members of society and placed near where they had died. Most barrows in Denmark are located in areas of productive farmland, evidence of the strong relationship between wealth and the control of agricultural resources (Figure 11.43). These high, circular mounds were often placed dramatically on the horizon to emphasize the importance of the buried individuals. Lines of barrows follow the high points on the landscape, along old roads and routes of movement. The distribution of thousands of such barrows in Denmark provides some information on the use of the landscape and
_ _
546
Chapter 11 Prehistoric Europe
the organization of early Bronze Age society. The amount of metal, bronze, and gold in these burials provides some indication of the wealth of the deceased individuals as well. All bronze and gold in southern Scandinavia had to be imported because the ores are not indigenous. Gold was more valuable; only 1 g (0.04 oz) of gold is found for every 1000 g (2.2 lb) of bronze. There are pronounced differences in buried wealth between the sexes and between individuals. Male graves are more frequent than female graves, and they contain more wealth. Some individuals were buried with a great deal of bronze and gold, and some without any, suggesting that social differentiation was pronounced in this region. An elite segment of the population must have controlled most of the resources as well as the trade.
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 547
Site
:
Vix A princess burial from the Iron Age
Iron has a melting temperature of over 3000°C (5400°F), and sophisticated furnaces and smelting techniques are required for reducing the ore. Iron making was discovered in Turkey shortly before 2000 B.C. The technology was probably a well-guarded secret for some time, to gain military advantage—iron can cut bronze. The new metal was used initially to make stronger, more durable weapons and later for making more practical tools and equipment. Iron came to Europe around the beginning of the first millennium B.C., slightly earlier in the eastern Mediterranean and slightly later in the northwestern part of the continent. The Iron Age in western Europe, the time of the Celtic tribes, can be divided into two phases: Hallstatt and La Tène. Hallstatt, the earlier period, approximately 800–500 B.C., was centered
in Austria, southern Germany, and the Czech Republic. During the early Iron Age, salt and iron mines in these regions led to economic boom times. The La Tène period followed Hallstatt. The major concentrations of sites from this period are found in eastern France, Switzerland, southern Germany, and the Czech Republic. The Celtic Iron Age came to an end in most of western Europe around 50 B.C. with the Roman conquest, led by Julius Caesar. The emperor Claudius conquered most of England shortly thereafter. Remnants of the Celtic traditions continued in Ireland for centuries, however, largely untouched by the Roman Empire. Germanic tribes in central and northern Europe also remained in the Iron Age on the fringes of the Roman Empire for another 1000 years, essentially until the arrival of Christianity.
Figure 11.44 Part of the decoration of the Gundestrup cauldron from Denmark, a classic example of Celtic art. This large silver bowl was found in pieces buried in a bog, perhaps as a sacrifice.
_ Site Vix
547
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Figure 11.45 The ground plan of the princess burial at Vix. The wagon bed in the center was used as a burial platform, and the wheels of the wagon were placed along the wall to the right. An enormous bronze krater (storage vessel) was placed in the upper-left corner of the tomb, along with a number of other vessels.
Page 548
Silver bowl
Bronze krater
Skull
Greek painted cup
Bronze bowls
Wagon bed
Brooches
Bronze bowls
Wagon wheels
0
y
Why did the introduction of bronze cause such dramatic changes in European societies?
_ _
548
A distinctive Celtic art style was practiced throughout western Europe during the pre-Roman Iron Age (Figure 11.44). Both Hallstatt and La Tène are defined primarily by styles of artistic depiction and decoration and by types of pottery. Designs from the La Tène period are both flamboyant and hypnotic. Complex patterns of concentric circles, spirals and meanders, and a variety of bird and animal figures appear on metal and ceramic objects. Disembodied heads with almond-shaped eyes and fierce mustaches; long, fanciful horse heads; and willowy statues of women characterize the tradition. Weapons, tools, jewelry, and everyday equipment were ornamented with this distinctive art. This art style, along with certain religious practices and beliefs, was shared by several distinct societies in
Chapter 11 Prehistoric Europe
1m
western Europe. These groups were headed by strong leaders and organized along similar lines. Some areas had elected magistrates. Julius Caesar, who encountered these groups in battle and in negotiations, wrote of their social structure, describing three major groups below the king: an aristocratic class of warriors and priests, the common people, and slaves—a stratified society. The tombs of the elite of Celtic society are among the best-known finds from this period. The grave at Vix (pronounced “vicks”) was excavated in 1953 by René Joffroy at the foot of an Iron Age hillfort (oppidum) known as Mont Lassois in northern France. A princess had been buried there beneath an earthen mound, along the headwaters of the Seine River, around 500 B.C. A large square chamber was dug into
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
the ground and lined with wooden planks. The body of a 35-year-old woman was placed on the bed of a ceremonial cart or hearse in the center of the grave. The wheels of the cart were removed and stacked against the sides of the tomb (Figure 11.45). Her body was covered with a leather blanket, and a bronze-headed staff was placed across her body. She was buried with a heavy gold collar of foreign manufacture (Figure 11.46), and wore Baltic amber beads, locally made bronze brooches, and other pins and jewelry. The tomb was filled with a wealth of exotic funerary offerings. A huge bronze krater, more than 1.5 m (5 ft) high, weighing 208 kg (450 lb), with a volume of more than 1000 liters (280 gal), was set in one corner of the grave (Figure 11.47). It was crafted in a Greek bronze workshop, probably located in southern Italy. The krater was designed to be dismantled for transport, and the assembly instructions were labeled with Greek letters. The design around its rim depicts Greek warriors and horse-drawn chariots. A number of bronze, silver, and gold bowls were placed alongside the krater, including a bronze flagon from Italy and two painted pottery cups from Greece. The origins of the materials found in the princess burial at Vix indicate the extensive trade taking place in Europe at that time and emphasize the importance of the major rivers as transportation routes. The Rhône, the largest river in Europe that joins the Mediterranean, provided a major corridor from the south into temperate Europe. The modern city of Marseilles, at the mouth of the Rhône River in southern France, was originally a Greek colony called Massilia, founded in 600 B.C. Wine from the Aegean was sent to Massilia for shipment into western Europe. Various goods and materials from the interior of the continent were exported through this colony to the eastern Mediterranean. Fragments of a garment of Chinese silk, found in a Hallstatt tomb in Germany, are one example of the extent of trade at this time. The quantity of materials flowing through Marseilles must have been remarkable.
Page 549
The fortress of Mont Lassois, directly above the grave at Vix, was one of the major political and commercial centers of late Hallstatt Europe. Mont Lassois dominates the upper Seine Valley at the spot where the river becomes navigable and flows west to the Atlantic. The Rhône and its tributaries lie just to the east. Thus, Mont Lassois is at a strategic point in the main route of commerce between the western Mediterranean and the Atlantic coast of western France, the English Channel, and the British Isles. Iron ore and forests for wood to make charcoal for reducing the ore were also available in the area. The richness of the offerings in the tomb of the princess at Vix, some 500 km (300 mi) north of Marseilles, suggests that the Greeks were giving gifts to the elite of Celtic society to obtain favorable trading status and secure commerce.
Figure 11.46 The imported gold collar found with the princess burial at Vix, 23 cm in diameter. The craftsmanship of the goldwork is remarkable.
krater A large metal vessel for mixing
and storing wine.
Figure 11.47 The enormous bronze krater found in the tomb at Vix. The vessel is 1.5 m high and weighs more than 200 kg. This wine container was probably made at the Greek colonies in southern Italy. Site Vix
549
_
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 550
Concept
<
The Bog People Iron Age cadavers from the peat: “creatures from the black lagoons”
In the peat cut, nearly seven feet down, lay a human figure in a crouched position, still half buried. A foot and a shoulder protruded, perfectly preserved but dark brown in color like the surrounding peat, which had dyed the skin. Carefully we removed more peat, and a bowed head came into view. As dusk fell, we saw in the fading light a man take shape before us. He was curled up, with legs drawn under him and arms bent, resting on his side as if asleep. His eyes were peacefully shut . . . his brows were furrowed and his mouth showed a slightly irritated quirk, as if he were not overpleased by this unexpected disturbance of his rest. —P. V. Glob (1970a)
Archaeologists rarely find flesh on the bones of the past. Much of what we know about our ancestors and their lives comes from the detective work of piecing together information from fragments of pots and tools, discarded bones and buried skeletons—the broken, forgotten, and hidden remnants of what human society once created. Unusual situations, especially in the case of human bodies preserved from the past, immediately capture our attention. These discoveries emphasize the fragile nature of the human condition. Among the most remarkable series of preserved prehistoric bodies are the “bog people” from northern Europe (Figure 11.48). Hundreds of individuals have been found in the peat bogs of northern Germany and southern Scandinavia, dating to the centuries around the birth of Christ. These bogs have marvelous preservative powers. The accumulation of peat and organic detritus that fills these swamps and mires contains tannic acid from the needles of coniferous trees. Tannic acid, used for tanning hides, is one essential factor in the preservation of the skin of the bog bodies. The second factor is the water-
Figure 11.48 Grauballe Man, found in a Danish bog in 1952. His throat had been cut, and he was placed naked in the waters of the Grauballe bog around 20 B.C. Many of the bog bodies known from northern Europe appear to have been individuals who were sacrificed or executed for crimes.
_ _
550
Chapter 11 Prehistoric Europe
logged condition of the bogs, which creates an environment where bacteria cannot break down soft tissues. The bog bodies are usually not the result of accidental drowning or disappearance. They are a curious consequence of the beliefs and practices of early Iron Age society. These people viewed the bogs as sacred places where sacrifices were made to the gods. Long braids of human hair were cut off and thrown into the bogs as one form of offering. Weapons and jewelry of bronze, iron, and gold were also placed in the mire to appease the gods. Individuals were intentionally sacrificed to the sacred bogs as well. Lindlow Man, found in an English bog in 1984, was bludgeoned with an axe and strangled, and finally his throat was cut. One of the best-known examples is Tollund Man, exhumed from the bogs of Denmark by P. V. Glob, of the Danish National Museum (Figure 11.49). Tollund Man was placed in the murky waters of a peat bog in central Denmark almost 2000 years ago. As Glob described it, the body was very well preserved; even his eyelashes and his beard were readily visible. Tollund
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 551
Figure 11.49 The head of Tollund Man, a bog person from Iron Age Denmark. The man was sacrificially executed and put in a bog. The corpse was preserved in the deposits there for almost 2000 years. The man was found with only a hat, a leather belt, and the leather garrote that had been used to strangle him.
Man had been strangled; a thin leather garrote was still tightly wound around his neck. He was naked except for a small leather cap and a belt around his waist. Tollund Man was well groomed when he died, and his hands were soft and uncalloused, not used for hard
labor. His last meal was a gruel of many kinds of seeds and grains. He died in the spring, and his execution may have been related to the rituals of Iron Age society that required sacrifice and offerings for the resurrection of the year and the bounty of the earth.
_ Concept The Bog People
551
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 552
Site
:
Maiden Castle The end of the Celtic Iron Age The largest settlements from the Iron Age in Europe are defended hilltops, found throughout southern Germany, France, and Britain. Mont Lassois above the Vix burial is one example, and there are many others across central and western Europe. These fortresses and their inhabitants were formidable. The Romans frequently remarked on the strength of the Gaelic walls that surrounded these hilltops. Roman writers reported on the ferocity of the Celtic warriors, describing the bold gaesatae, spearmen who went naked into battle. Roman statuary commemorated the valor of the Celts as fair-haired warriors with drooping mustaches. In a famous sculpture, a defiant, yet defeated, Celt puts himself to the sword rather than surrender to the Romans. That the Romans should so admire the character of their boisterous and aggressive enemy is itself a remarkable statement about the Celts. Hundreds of Iron Age hillforts are found in Britain and Ireland in various sizes, from less than an acre to many hectares. These hillforts served as both
Figure 11.50 Maiden Castle in Dorset, England, showing the fortifications of multiple walls and ditches. This high hill has archaeological remains from the Neolithic through the Roman period.
_ _
552
Chapter 11 Prehistoric Europe
population centers and retreats during the Iron Age, distinguished by the fortifications that surrounded them. Maiden Castle is one of the largest hillforts in Britain; in its final form, it enclosed an area of almost 18 ha (45 acres). Situated atop a high, saddleback hill in the downs of Dorset, near the southern coast of England, Maiden Castle dominated the Iron Age Dorset landscape (Figure 11.50). Sir Mortimer Wheeler excavated parts of the site from 1934 through 1937, and more recent excavations were undertaken during the 1980s. Wheeler was a major figure in British archaeology, highly respected for both the quality of his work and his concern with the education of archaeologists. After his college studies, Wheeler served as an officer in World War I, the commander of a battery in the Royal Artillery. Returning at the end of the war to head the archaeology section of the National Museum of Wales, in 1925 he became director of the London Museum. His excavations at Maiden Castle were intended both to reveal
pri31057_ch11.pp506-559.qxd
10/1/09
W. Knoll
12:14 PM
Neolithic settlement
Page 553
Neolithic lo
ng mound
E. Knoll Round barrow
Working floor Neolithic Phase I
Figure 11.51 Wheeler’s interpretation of the construction phases at Maiden Castle, from the Neolithic through the Iron Age.
Pits
Neolithic Phase II
Iron Age (A) Phase I
Iron Age (A) Phase II
Iron Age (Early B) Phase III
Iron Age (Late B) Phase IV
the history of the site and to train new students in archaeology. Wheeler’s forte was the stratigraphic section. This vertical wall, or profile, of excavated trenches and squares provides information essential to understanding the history of a site. Recording these sections in photographs and drawings creates a permanent record of the sequence of events for the archaeologist to interpret. Wheeler’s precise excavations and recording methods have revealed much about the prehistory of Maiden Castle, from the Neolithic to the Roman conquest of Britain. The hilltop was first used around 3700 B.C. as a Neolithic camp; an enormous barrow was erected here during the same period, overlooking the adjacent valley (Figure 11.51). This extraordinary Neolithic barrow ran for over 550 m (1800 ft) across the hilltop, marked by two parallel ditches, each 5 m (16 ft) wide, 15 m (50 ft) apart. Maiden Castle was the center of an elaborate landscape of henge monu-
ments, timber settings, and causewayed camps, comparable to other major Neolithic centers in the British Isles. During the Iron Age around 500 B.C., fortifications were first constructed around a growing market center on the hilltop. The first fort at Maiden Castle was univallate, having one wall, and enclosed 6 ha (15 acres). Large hillforts such as Maiden Castle appear to have had administrative, religious, economic, and residential functions in a proto-urban context, an early form of the walled town. There are several distinct areas within Maiden Castle. An Iron Age shrine or temple was uncovered on the most prominent part of the hilltop. This structure was later replaced by a Roman temple, documenting the continuity of religious significance and focus. The hilltop served as a residence as well as a fortified storehouse for the agricultural surplus from the rich farmlands. Residential structures dot the hilltop; evidence of dense occupation suggests that as many as 4000
With so much trade and interaction, why was warfare endemic in much of Europe’s past?
y
stratigraphic section The excavation
of trenches and squares across manmade layers to expose a cross section of the deposits and reveal the sequence and methods of construction.
Site Maiden Castle
553
_
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 554
Figure 11.52 A Roman catapult arrowhead, found in the back of the skeleton of one of the defenders of Maiden Castle, poignantly marks the end of the Celtic Iron Age in western Europe and its replacement by the Romans.
y
_ _
How were the Romans able to conquer many of Europe’s Iron Age inhabitants? Why not the Germanic tribes?
multivallate A term describing complex defenses of multiple ditches and ramparts at large Iron Age hillforts.
554
people lived there. Large pits were dug into the chalk, as much as 3 m (10 ft) deep for storage, water reservoirs, and other purposes. Excavations also revealed the parallel ruts of wagon wheels crossing the defended interior; the standard wheel-to-wheel distance in the later Iron Age was 1.46 m (4.8 ft). More recent excavations have uncovered numerous structures and an enormous quantity of refuse in the form of pottery, bronze and iron artifacts, and animal bones. Most of the animal bones were from sheep, and several had signs of malnutrition. In 50 B.C., shortly before the first Roman invasion of Britain, the fortifications were enormously expanded, and the enclosed area inside the hillfort tripled in size. The final set of fortifications consisted of a set of three enormous concentric banks and two ditches enclosing almost 18 ha (45 acres). A fourth bank was built at the south end for additional protection. The term multivallate is used to describe large hillforts with complex defenses of multiple ditches and ramparts. Some of the walls at Maiden Castle stand as high as 20 m (65 ft) today. These walls had a 4-m (13-ft)-high foundation of stone and a massive wooden palisade on top. A narrow, serpentine, and dangerous path guarded the entrance at the eastern end of the hilltop. At the entrance were wooden gates 4 m (13 ft) wide. Some 22,000 slingstones were found in caches near the walls of Maiden Castle, representing one of the important
Chapter 11 Prehistoric Europe
defensive weapons in use during that period. In spite of its massive defenses against attack, Maiden Castle fell to the Roman legions and their siege artillery in A.D. 43 (Figure 11.52). Wheeler’s depiction of the final hours of the defenders of Maiden Castle is a masterpiece of archaeological interpretation and prose: Before the close fighting began, the regiment of catapults or ballistae, which habitually accompanied a legion on campaign, put down a barrage across the gateway, causing casualties at the outset. Following the barrage, the Roman infantry advanced up the slope, cutting its way from rampart to rampart, tower to tower. In the innermost bay of the entrance, a number of huts had been recently built; these were now set alight, and under the rising clouds of smoke the gates were stormed. But resistance had been obstinate and the attack was pushed home with every savagery. The scene became that of a massacre in which the wounded were not spared. Finally, the gates were demolished and the stone walls which flank them reduced to the lowly and ruinous condition in which we found them, nineteen centuries later. (Wheeler, 1943, p. 47)
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 555
Images and Ideas Lessons from Prehistoric Europe The growth of European civilization
T
he story of European prehistory is one of spectacular growth and change. Because of a long history of archaeological research and the richness of prehistoric remains, we know more about the past of Europe than of almost any other place. The prehistory of Europe is, in fact, one model of the evolution of society, from small bands of hunter-gatherers in the late Pleistocene and early Holocene, through the first hierarchical societies of the early Neolithic, to the powerful states of the Bronze and Iron Ages (Figure 11.53). The introduction of agriculture resulted in major changes in European society over the six millennia before Christ. The initial spread of farming was slow, taking place over a period of roughly 3000 years. The first farming communities in Europe are known from Greece and the Aegean area and date to around 6500 B.C. By 3000 B.C., farming had penetrated to the coasts of western Europe and reached its climatic limits in northern Scandinavia and Britain. By that time, virtually all the cultivable areas of Europe were inhabited by farming populations, and huntergatherers continued only in the most northerly, marginal areas. In the third millennium B.C., these initial Neolithic societies became more regionalized, following idiosyncratic trends but participating widely in increasing warfare, wealth, and trade. The transition to farming appears to have followed a common pattern in much of Europe, as domesticated plants and animals were slowly incorporated into a subsistence pattern based largely on wild foods. In northern Europe, at least, the shift from a reliance on wild foods to a predominance of domesticates took several thousand years. The spread of agriculture was accompanied by significant changes in social relations. New religious beliefs were spreading at this time as well. One manifestation is the appearance of megalithic tombs along the Atlantic coastline of Europe, whose construction represented enormous efforts. The monuments themselves would have been dramatic statements of the authority of the individuals buried within them. This inequality was also represented in the appearance of rare objects such as copper and gold. Apart from the elite control of highly valued materials, economies were probably based on barter and reciprocal exchange. This trend toward hierarchical organization in the Neolithic accelerated with the introduction of bronze, around 3000 B.C. The third and second millennia B.C. witnessed the rise of early states and empires in Mesopotamia, the Nile Valley, and the Aegean. Before the literate civilizations of classical Greece and Rome, the Aegean area witnessed the emergence of state-level societies on Crete and the Greek mainland. The Minoan palaces and Mycenaean citadels were urban centers of these societies and the focal points of industry, commerce, religion, military power, and central accumulation. During the third millennium B.C. in Europe, several major changes in technology and transportation occurred, including the manufacture of bronze weapons and jewelry, the widespread adoption of the plow and the wheel, and the use of draft animals such as oxen and horses to pull carts and chariots. Weapons and martial motifs dominated the symbols of status in European society during this time. Local warfare and conflict contrasted strongly with the long-distance patterns of trade, commerce, and cooperation. Textiles of wool and linen, metal dress pins, and amber and gold necklaces characterized the clothing of the elite. A bronze razor (for shaving) was a means to alter one’s appearance as a mark of status and power. Certainly by this time, rituals were scheduled according to regular times of the year. The solar “computer” of Stonehenge and the myriad stone Images and Ideas Lessons from Prehistoric Europe
< ( : < V : y < p : W 555
_
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 556
Taxation
Figure 11.53 The evolution of European society superimposed on the chart of social evolution. The major changes over time include increased size and density of population and more hierarchical organization. Warfare, trade, and social inequality grew in intensity from the Neolithic through the Iron Age.
Rome
Military draft Bureaucracy
Greece
True law
Classical
Kingship Stratification Full-time craft specialization
Mycenae
Separate marriage networks for elites and commoners Time
Hereditary leadership
Knossos
Maiden Castle
Central accumulation
Vix
Ranked descent groups
Varna
Regular calendrically scheduled rituals
Borum Eshøj
Pan-tribal associations, fraternal orders
Stonehenge
Unranked descent groups Reciprocal or barter economy Unscheduled ad hoc ritual
Charavines
Ephemeral, informal leadership
Franchthi Cave
Egalitarian status Local group autonomy
and wood circles of northwestern Europe clearly functioned in this manner. North of the Alps, political and social organization seems to have involved alliances of ephemeral chiefs who cooperated in the exchange of exotic materials and warfare. The economy and social relations were intertwined. As wealth accumulated and increased during the Bronze Age, there was a marked trend toward fortification. The most notable features of the Bronze Age landscape, aside from exceptional monuments such as Stonehenge, are large barrows—the tombs of the elite. There was a dramatic, intentional contrast between the elevated graves of the nobility and the flat cemeteries of the commoners. The introduction of iron after 1000 B.C. brought new tools and weapons to Europe. Iron produced harder and sharper implements and was in great demand. In contrast to copper and tin, whose sources were limited to only a few areas in Europe, iron ores in a variety of forms were found in many places across the continent. Veins of iron ore were exploited in Iberia, Britain, the Alps, the Carpathian Mountains, and elsewhere. Bog iron was exploited in northern Europe, and carbonate sources in other areas enabled local groups to obtain the raw materials necessary for producing this important material. At the same time, the collapse of the dominant civilizations of the Aegean area around 1000 B.C. changed the flow of raw materials and finished products across Europe. However, new centers in the Mediterranean, first in the classic
_ _
556
Chapter 11 Prehistoric Europe
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 557
civilization of Greece and later in the Roman Empire, brought history to Europe as these powerful empires emerged. Again, we can observe important and dramatic differences between the “classic” areas of the Mediterranean and the northern parts of “barbarian” Europe. Several events marked the beginning of the La Tène period of the Iron Age, between 500 B.C. and the birth of Christ. The collapse of trade with the Mediterranean, rapid population increase, and conflict led to expansion and a series of migrations to the south and east of the Celtic heartland. The Celts, or Gauls, moved into northern Italy and the Po River Valley, and farther east into Romania and eventually Anatolia in Asia Minor. Around 390 B.C., the Gauls sacked Rome. With the decline of long-distance trade, local craftsmanship improved, and a spectacular, powerful style of art and decoration predominated among the Celts. These crafts, along with many other aspects of society, were concentrated in towns in the Celtic part of Europe. These towns most often occupied the oppida, the heavily walled and fortified centers of political and economic power. Maiden Castle is just one of many examples of these Celtic towns that fell to the Romans as the empire expanded. A very different sequence of events characterized the Mediterranean. By 800 B.C., the effects of growing population and urbanization led to the emergence of Greek city-states as major powers. Conquest and continuing population expansion led to Greek colonization of several areas to the east and west of the homeland. Colonies appeared in Sicily, southern Italy, southern France, and Spain, because the eastern Mediterranean could not contain the growing numbers of Greeks, Etruscans, and Carthaginians. Local demand for Greek goods was high in these areas. At a site in Spain, for example, 1400 Greek pottery vessels were found in a single, small excavation. At another site in Spain, a single grave contained gold jewelry that included a seal from Syria and an earring from Egypt. The Greeks, and later the Romans, had encoded systems of law, military drafts, and elaborate bureaucracies, characteristics of high-level state societies. As the Greece of Homer and Socrates declined in power, Rome began to build its empire. The Macedonians under Philip, and later Alexander the Great (356–323 B.C.), had turned to the east, setting out against Asia. Trade and colonization during the earlier half of the Iron Age, the first millennium B.C., changed to warfare after 400 B.C. as the dynamics of power shifted to the Italian peninsula. By 290 B.C., the Romans had spread into most of Italy, and by 140 B.C., they had eliminated their major rivals at Corinth in Greece and Carthage in North Africa. Rome controlled the Mediterranean and turned toward the interior of Europe. Roman trade and colonization at first focused on Spain as an important source of raw materials. Soon, however, trade routes and the wealth of northern Europe captured the attention of the Romans. Julius Caesar led the legions against the Gauls who occupied France, the British Isles, and parts of central and eastern Europe. From 59 to 51 B.C., Caesar pushed the boundaries of the Roman Empire across France to the Rhine River, largely destroying Celtic society in the process. In one instance, following the conquest of the town of Cenabrum, the Romans killed all but 300 of the 40,000 men, women, and children taking refuge in the fortress. A few decades later, the empire extended across the Channel into much of England. By the birth of Christ, Rome controlled virtually all of southern Europe. A few pockets of Celts remained in Ireland and Scotland and on some Atlantic and Channel islands. North of the Rhine, however, the Germanic tribes resisted and eventually outlasted the Roman Empire. The Germanic Iron Age followed the Roman period and extended from the birth of Christ through most of the first millennium A.D. In this area of Germany, Scandinavia, and eastern Europe, warrior societies continued and flourished (Figure 11.54). These societies consisted of small, yet highly
Figure 11.54 A copy of a twelfthcentury A.D. Norse game piece from the Isle of Lewis, British Isles.
Images and Ideas Lessons from Prehistoric Europe
_
557
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 558
productive, agricultural communities. Society was largely egalitarian, and authority was vested in individuals of demonstrated strength and wit. Through time, these communities grew larger and more powerful, gradually forming alliances to create larger polities. While the Roman conquest effectively ended prehistory in southern Europe and Britain, the spread of Christianity marked the beginning of the historic period in most of Europe north of the Alps. The spread of Christianity was followed by the Dark Ages, the Renaissance, the industrial era, and eventually the present.
p
DISCUSSION QUESTIONS 1. What are the characteristics of the three major periods of European prehistory?
w w w. m h h e . c o m / p r i c e i p 6 e
For more review material and study questions, see the self-quizzes on your online learning center.
2. Describe the introduction of agriculture into Europe in three paragraphs. 3. What was the impact of the rise of the Aegean civilizations in the rest of Europe? 4. Discuss your impressions of life in the Bronze Age from the evidence of tombs. 5. Who were the Celts and what was their role in European prehistory? 6. The prehistory of Europe is a study in contrasts: trade and warfare, elite and common, town and village, stone and metal. Describe one such contrast with an example.
z
_ _
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e.
Bradley, R. 1998. The significance of monuments. London: Routledge. A refreshing review and innovative perspective on the burial monuments of western Europe. Chadwick, J. 1976. The Mycenaean world. Cambridge: Cambridge University Press. A classic treatise on the Mycenaeans, with particular emphasis on the written language. Chippendale, C. 1983. Stonehenge complete. Ithaca, NY: Cornell University Press. A well-illustrated compendium on the evolution of Stonehenge and our understanding of it. Cunliffe, B. 2008. Europe between the oceans: 9000 B.C.– A.D. 1000. New Haven, CT: Yale University Press. A very readable synthesis of the last 10,000 years of European prehistory. Cunliffe, B., ed. 1994. The Oxford illustrated prehistory of Europe. Oxford: Oxford University Press. A recent synthesis of European prehistory, written by various experts in the field. Foster, J. 2003. Life and death in the Iron Age. Oxford: Ashmolean Museum. A popular account of the Iron Age and its archaeology. French, E. 2002. Mycenae: Agamemnon’s capital. Stroud, UK: Tempus. A discussion of Mycenae and the Mycenaeans by a leading authority. Harding, A. F. 2000. European societies in the Bronze Age. Cambridge: Cambridge University Press. A comprehensive, authoritative, and up-to-date account of Bronze Age societies in Europe.
558
Chapter 11 Prehistoric Europe
Jones, A., ed. 2008. Prehistoric Europe: Theory and practice. Oxford: Blackwell. New perspectives on selected parts of European archaeology. Kristiansen, K. 2000. Europe before history. Cambridge: Cambridge University Press. A theoretical consideration of the later prehistory of Europe, largely north of the Alps. Milisauskas, S., ed. 2002. European prehistory, a survey. New York: Springer. A survey of prehistoric Europe by a group of leading experts. Perlès, C. 2001. The Early Neolithic in Greece. Cambridge: Cambridge University Press. A new classic on the beginnings of agriculture in Europe. Powell, T. G. E. 1980. The Celts. London: Thames & Hudson. An updated classic on the Celtic peoples of Europe and their art. Price, T. D. 2000. Europe’s first farmers. Cambridge: Cambridge University Press. A series of regional papers on the introduction of agriculture in Europe. Spindler, K. 1994. The man in the ice: The discovery of a 5,000-year-old body reveals the secrets of the stone age. Bielefeld, Germany: C. Bertelsmann Verlag. The archaeologist in charge reporting on the Iceman and results of his study. Whittle, A. 2003. The archaeology of people: Dimensions of Neolithic life. London: Routledge. A revised perspective on the European Neolithic.
pri31057_ch11.pp506-559.qxd
10/1/09
12:14 PM
Page 559
_
pri31057_ch12.pp560-575.qxd
C
_ _
10/1/09
H
A
8:30 AM
P
Page 560
12
T
E
R
T
W
E
Figure 12.1 Square Tower House ruin in Mesa Verde National Park, Colorado, c. A.D. 1250.
L
V
E
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 561
< The Past as ( Present and Future : Introduction < A V : y < p : W CHAPTER OUTLINE
The Past as Present and Future
Introduction The Past as Present and Future 561
rchaeology is beautiful. The Square Tower House ruin in Mesa Verde National Park in Colorado contains some of the largest and most beautiful of the cliff dwellings in the southwestern United States (Figure 12.1). Archaeology is also fragile; the past doesn’t live forever. Archaeology is worth protecting. To conclude Images of the Past, it seems appropriate to turn toward the present and the future, to examine what we have learned in the context of our lives today. The following pages bring together several diverse aspects of archaeology that don’t really fit in the review of major sites, finds, and concepts that has filled the preceding chapters. Under separate headings, we consider issues of the relevance of archaeology, archaeological heritage, and ethics. Although these topics are discussed separately, they are in fact closely related aspects of the past in the present. In case you are interested, there is also a section on careers in archaeology.
THE VALUE OF THE PAST 561 THE HERITAGE OF THE PAST 563
WHO OWNS THE PAST? 564 U.S. LEGISLATION AND ARCHAEOLOGY 567 Kennewick Man 568 ETHICS IN ARCHAEOLOGY 569
THE RESPONSIBLE ARCHAEOLOGIST 569
THE VALUE OF THE PAST
A very reasonable question concerns the relevance of archaeology. Why should we study archaeology? Why is it important to know about what happened in prehistory? One simple reason to learn about archaeology is to be able to participate in the contemporary fascination with the subject. Public opinion in the United States is very positive with regard to archaeology. A recent poll revealed that 76% of those questioned expressed a real interest in archaeology and 90% thought archaeology should be taught in school. Reasons for that interest were varied; learning about how people lived in the past, connecting the past to the present, and the thrill of discovery were mentioned. Archaeology is clearly of inherent interest to the general public. Stories appear almost daily in our newspapers and on television. The Internet is crowded with sites about archaeology and our human past. But the public’s fascination with archaeology stems from the inherent significance of the subject. There is, of course, an exciting mystery involved in unearthing treasures from the earth, but, more than that, archaeology tells us about ourselves and how we got to be the way we are. Ultimately, it is the lessons archaeology teaches about our past that convey its significance. The human condition is one that has changed and will change over time. To know our place and to have confidence about where we are going are essential ingredients in the success of our species.
561
CAREERS IN ARCHAEOLOGY 572
Images and Ideas The End 574
There is no present or future—only the past, happening over and over again— now. —Eugene O’Neill
w w w. m h h e . c o m / p r i c e i p 6 e
For preview material for this chapter, see the comprehensive chapter outline and chapter objectives on your online learning center.
_
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 562
Figure 12.2 The Rock of Cashel, Ireland, a fortress and religious center dating back to the fourth century A.D.
Study the past if you would define the future. —Confucius, 561–479 B.C.
_ _
562
Gathering and/or hunting constituted most of our human career. The roots of human behavior are to be found in our past as hunter-gatherers. The changes that took place in that period were both biological and cultural: primarily biological at the beginning, almost exclusively cultural toward the end. Our earliest hominid ancestors appeared 5–4 m.y.a. in Africa. The first groups of farmers have been found in western Asia around 11,000 years ago. During the intervening period of roughly 4 million years, our activities shifted from a peripatetic quest for food by small groups of a generic apelike ancestor to an elaborate pattern of intensive food collecting and storing by large groups of fully modern humans living in permanent communities. In fact, hunter-gatherers have continued to exist at the margins of modern society until very recently. Understanding the archaeology of the beginnings of humanness, the origins of fully modern behavior, and our background as foragers is essential to understanding ourselves and our place in nature. In a sense, our life in megapolistic societies with highly structured hierarchical organizations, highly advanced technologies, and extraordinary communication and exchange is too new to be understood without knowledge of the baseline behaviors from which it derived. Archaeology tells how humans, over thousands and millions of years, have survived and succeeded in the face of the difficult challenges of changing environments and competitive neighbors. We learn how past societies dealt with issues such as environmental change, overpopulation, and political competition. Maybe one of the most important lessons of archaeology is that we are only building on what has been done before. The role of archaeology ultimately is to describe the course of human development, to tell us about our origins. That knowledge can provide both pride and confidence in our species, along with a much greater awareness of our oneness. Of course, that may not seem so important in an age when recent polls indicate that the majority of Americans do not know that the earth travels around the sun every 365 days and continue to believe that humans and dinosaurs did battle in the past. But then again, maybe that is why it is so important. Archaeology has the opportunity to greatly extend the fundamental contributions that are made to human knowledge and society.
Chapter 12 The Past as Present and Future
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 563
Figure 12.3 Standing stone giant of Easter Island, South Pacific, dating to approximately A.D. 1500.
THE HERITAGE OF THE PAST But we also learn larger lessons from the past. The Spanish American philosopher Santayana wrote that those who cannot remember the past are condemned to repeat it. Certainly, there are lessons about the long span of our prehistory on earth and about the evolution of our species and our behavior. In the vastness of geological or archaeological time, the role of humans on the planet is minuscule indeed. Yet the impact of our species is immeasurable. Although relative newcomers to earth’s history, we have an obligation and a responsibility toward all that is around us. Humans have tremendous destructive power as well as creative abilities. Through looting, careless development, and the wanton destruction of archaeological resources, we even have the potential to eliminate our capacity to reconstruct and understand our own past. There is a note, rather a crescendo, of urgency in our quest for scientific understanding of our ancestral history. The archaeological record, under attack from growing populations and global economies, is rapidly disappearing in many areas. The growth of modern cities and transportation networks is covering the earth’s surface with concrete, asphalt, and mountains of trash from our consumer society. The industrialization of agriculture results in the deep plowing of sites, the mixing of layers of soil, and the lowering of the groundwater table, destroying the archaeological record. If we are to have any archaeology at all in the future, it is essential that the fundamental information on our past that remains buried in the ground be recorded and protected, before there remains nothing for us to study. This challenge should be a major focus of concern and action in the twenty-first century. There are a number of ways to help protect the past. One of them is simply to learn more and to tell others of the importance of our archaeological heritage. Public interest and awareness is needed to safeguard the past. A number of organizations of archaeologists and others are working to increase public awareness and support. In the United States, the Society for American Archaeology is notably active in this effort. Another step is to recognize that no one owns the past—it belongs to all of us; important archaeological sites and artifacts should not be the property of individuals (Figures 12.2 and 12.3). Dealers in antiquities and the market they supply are responsible for the looting of hundreds of archaeological sites each year. Enforcing existing legislation against the import and sale of antiquities is a large part of the battle to save the past.
That’s the thing about history, it’s over before you know it. —Garrison Keillor
_ The Heritage of the Past
563
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 564
TABLE 12.1 The Mission of UNESCO’s World Heritage Program • Encourage countries to sign the 1972 Convention and to ensure the protection of their natural and cultural heritage. • Encourage countries to nominate sites within their national territory for inclusion on the World Heritage List. • Encourage countries to set up reporting systems on the state of conservation of World Heritage sites and establish management plans. • Help countries safeguard World Heritage sites by providing technical assistance and professional training. • Provide emergency assistance for World Heritage sites in immediate danger. • Support activities that increase public awareness of World Heritage conservation. • Encourage participation of the local population in the preservation of their cultural and natural heritage. • Encourage international cooperation in conservation of cultural and natural heritage.
A number of organizations are actively working to protect archaeological sites. Two of the larger ones are the Archaeological Conservancy and the UNESCO World Heritage Site program. Both are worthy programs deserving support. The Archaeological Conservancy, established in 1980, is the only national nonprofit organization dedicated to acquiring and preserving remaining archaeological sites in the United States. Since its founding in 1980, the conservancy, based in Albuquerque, New Mexico, has acquired more than 245 sites in 37 states across America. These range in size from a few acres to more than 1000 acres and include the earliest habitation sites in North America, a nineteenth-century frontier army post, and remains from nearly every major cultural period in between. The United Nations Educational, Scientific, and Cultural Organization (UNESCO) was chartered in 1972 to encourage the identification, protection, and preservation of cultural and natural heritage around the world considered to be of outstanding value. “Cultural heritage” refers to monuments, buildings, and sites with historical, aesthetic, archaeological, scientific, ethnological, or anthropological value. “Natural heritage” refers to outstanding physical, biological, and geological formations; habitats of threatened species of animals and plants; and areas with scientific, conservation, or aesthetic value. The mission of this program is described in Table 12.1. To date, 830 sites have been enscribed in the program, some of which are listed in Table 12.2.
WHO OWNS THE PAST? One issue facing archaeology today is growing concern about who owns the past. This issue involves both national and individual ownership. When archaeology was just beginning and European empires controlled much of the known world, soldiers and merchants from the conquering nations returned home with many treasures, both ancient and modern, from the countries they had visited. For example, Napoleon returned from his invasion of Egypt with a number of marvelous ancient Egyptian artifacts and monuments. A well-known example involving ownership of the past concerns the Elgin marbles. These statues have a long and storied history. After the Greeks defeated the Persians in 479 B.C., the Greeks returned to Athens to resurrect their city, ruined during the war. Their leader, Pericles, rebuilt the city as an artistic and cultural center with many wonderful structures—including the Parthenon—during the 30 years of his rule. The Parthenon took 15 years to construct and was finally dedicated in 432 B.C. as the temple of the goddess Athena, sitting on the Acropolis above Athens as the pinnacle of Greek power and creativity.
_ _
564
Chapter 12 The Past as Present and Future
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 565
TABLE 12.2 A Partial List of World Heritage Archaeological Sites* Site
Location
Prehispanic city of Chichén Itzá Stonehenge, Avebury, and associated sites Angkor Prehispanic city and national park of Palenque Heart of Neolithic Orkney Imperial tombs of the Ming and Qing dynasties The Great Wall Acropolis, Athens Archaeological sites of Mycenae and Tiryns Prehispanic city of Teotihuacan Archaeological park and ruins of Quirigua Decorated grottoes of the Vézère Valley Archaeological site of Vergina Chan Chan archaeological zone Prehistoric rock-art sites in the Côa Valley Altamira Cave Neolithic flint mines at Spiennes (Mons) Thracian tomb of Kazanlak Pont du Gard (Roman aqueduct) Archaeological ensemble of the Bend of the Boyne Petra Lines and geoglyphs of Nasca and Pampas de Jumana Rock carvings in Tanum Rapa Nui National Park Memphis and its necropolis—the pyramid fields from Giza to Dahshur Persepolis Great Zimbabwe National Monument Archaeological ruins at Mohenjo-daro Lower valley of the Awash Fatehpur Sikri Hadrian’s Wall Archaeological site of Troy San Agustín Archeological Park Jelling mounds, runic stones, and church Rock drawings of Alta Chavín archaeological site Koch’ang, Hwasun, and Kanghwa dolmen sites Rock-art of the Mediterranean Basin on the Iberian Peninsula Ban Chiang archaeological site Mesa Verde Chaco Culture National Historical Park Tiwanaku: spiritual and political center of the Tiwanaku Joya de Ceren archaeological site Lower valley of the Omo Archaeological sites of Bat, Al-Khutm, and Al-Ayn City of Cuzco Birka and Hovgården Cahokia Mounds State Historic Site Archaeological areas of Pompei, Herculaneum, and Torre Annunziata Archaeological site of Atapuerca Peking Man site at Zhoukoudian Roman theater and its surroundings and the “Triumphal Arch” of Orange Sangiran early man site Fossil hominid sites of Sterkfontein, Swartkrans, Kromdraai, and environs Megalithic temples of Malta Maya site of Copan L’Anse aux Meadows National Historic Site Head-Smashed-In Buffalo Jump
Mexico United Kingdom Cambodia Mexico United Kingdom China China Greece Greece Mexico Guatemala France Greece Peru Portugal Spain Belgium Bulgaria France Ireland Jordan Peru Sweden Chile Egypt Iran Zimbabwe Pakistan Ethiopia India United Kingdom Turkey Colombia Denmark Norway Peru Republic of Korea Spain Thailand United States United States Bolivia El Salvador Ethiopia Oman Peru Sweden United States Italy Spain China France Indonesia South Africa Malta Honduras Canada Canada
*The sites listed here constitute only a part of the 830 places of great historical, archaeological, cultural, or natural importance protected by United Nations Convention. The list is growing, but not quickly enough.
_
pri31057_ch12.pp560-575.qxd
10/1/09
Figure 12.4 The Elgin marbles at the British Museum.
_ _
566
8:30 AM
Page 566
Atop the Parthenon, just under the roofline, was a frieze of more than 400 human and 200 animal marble statues. The frieze has a maximum height of more than 2 m (about 7 ft) and a total length of 160 m (about 500 ft) around the four sides of the temple. One of the marvels of antiquity, the frieze depicts the most important annual festival of ancient Athens. During the six-day celebrations, various musical, literary, and athletic competitions were held, including beauty contests, track-and-field events, horse and chariot races, and regattas. In 1801, Thomas Bruce, Lord Elgin and the British ambassador to Constantinople (Turkey), obtained permission to remove large portions of the Parthenon frieze. From Athens, the marble statues of the Parthenon were shipped to the British Museum in London, where they remain today, known as the Elgin marbles (Figure 12.4). The Greek government has been petitioning the British government for many years for the return of these treasures, but to no avail. The United Nations, the British populace, and world opinion support the Greek claim, but British authorities counter with statements about setting a precedent for the return of other museum treasures around the world. The British also argue that the marbles were purchased legitimately from the Turks who controlled Athens at that time, that the marbles were removed in order to save them from destruction, and that the Greeks did not care then about the marbles. More recently, the buildup of corrosive air pollution in Athens has been added to the arguments against returning the marbles. One moral of such cases is that there are rarely easy answers to questions about the ownership and/or return of cultural heritage. To reiterate: No one owns the past; it belongs to all of us. Important archaeological sites and artifacts should not be the property of individuals. Archaeological materials are valuable cultural resources and part of our inheritance. It is essential that everyone understands their significance and works toward their care and curation. Trading in antiquities or disturbing archaeological sites without permission is illegal in most states and punishable by fines and/or imprisonment. Dealers in antiquities and the market they supply are ultimately responsible for the looting of hundreds of archaeological sites each year. Enforcing existing legislation against the import and sale of antiquities is a large part of the battle to save the past.
Chapter 12 The Past as Present and Future
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 567
TABLE 12.3 Major Stipulations of the Native American Graves Protection and Repatriation Act (NAGPRA) of 1990 • Federal agencies and museums must identify cultural items in their collections that are subject to NAGPRA and prepare inventories and summaries of the items. • Federal agencies and museums must consult with lineal descendants, Indian tribes, and Native Hawaiian organizations regarding the identification and cultural affiliation of the cultural items listed in their NAGPRA inventories and summaries. • Federal agencies and museums must send notices to lineal descendants, Indian tribes, and Native Hawaiian organizations describing cultural items and lineal descendancy or cultural affiliation and stating that the cultural items may be repatriated. The law requires the secretary of the interior to publish these notices in the Federal Register.
U.S. LEGISLATION AND ARCHAEOLOGY The first U.S. government legislation to protect the past was signed by President Theodore Roosevelt in 1906. The Antiquities Act described the study of archaeology as a scientific undertaking and protected 167 million acres of cultural and natural significance. The act gave the president unilateral power to create national monuments on government property; the act also enumerated punishment for individuals caught looting or damaging national monuments and listed qualifications for individuals permitted to conduct research. This act was the precursor to a series of subsequent legislative decisions. The U.S. Congress reiterated the importance of the historic and prehistoric past of the country in 1966 with the passage of the Natural Historic Preservation Act. This act recognized the importance of cultural resources and established the National Register of Historic Places and the legal framework for the protection of cultural resources. Because of growing public interest in the past and a substantial increase in looting and the illicit trade in antiquities, Congress passed the Archaeological Resources Protection Act in 1979 to further help preserve archaeological materials on public and Indian lands. This law defined archaeological resources to be any remains of past human life or activities that are of archaeological interest and are at least 100 years old; encouraged cooperation between groups and individuals in possession of archaeological resources from public or Indian lands with special permit and disposition rules for the protection of archaeological resources on Indian lands; provided that information regarding the nature and location of archaeological resources remain confidential (to prevent looting and disturbance); and established civil and criminal penalties, including forfeiture of vehicles, fines of up to $100,000, and imprisonment of up to 5 years for second violations for the unauthorized appropriation, alteration, exchange, or other handling of archaeological resources, with rewards for furnishing information about such unauthorized activities. An even more sensitive and contentious issue involves the disposition of ancient human skeletal remains. North American archaeologists have been excavating burials, along with other cultural materials, for many years. By some estimates, more than 100,000 Native American graves have been excavated in the United States and the skeletons placed in museums. The Smithsonian Institution in Washington, DC, for example, houses the remains of more than 16,000 Native Americans collected over the years. Because of growing concern about the excavation of native remains and the desire of native peoples to rebury these remains, Congress passed the Native American Graves Protection and Repatriation Act (NAGPRA) in 1990. NAGPRA provides a mechanism for museums and federal agencies to return certain Native American cultural materials—human remains, funerary and sacred artifacts, and objects of cultural patrimony—to lineal descendants, _ U.S. Legislation and Archaeology
567
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 568
culturally affiliated Indian tribes, and Native Hawaiian organizations. The major features of NAGPRA are listed in Table 12.3. Items requested for return must be repatriated to a lineal descendant or related group. Several different lines of evidence are required to determine cultural affiliation, including geographic, biological, archaeological, linguistic, and anthropological evidence. The law also forbids trafficking in Native American cultural or human material and establishes procedures for notification of and consultation with tribes for planned excavation or accidental discovery of cultural materials on tribal property. Because of NAGPRA and similar legislation, the Smithsonian has repatriated more than 4000 sets of human remains to native tribes for reburial, out of a total collection of some 16,000, since 1984. The repatriation of human skeletal remains is not always a clear-cut issue, however, and this legislation has created several controversial situations, such as the case of Kennewick Man.
Kennewick Man
Figure 12.5 The cover of Time in March 2006, continuing the saga of Kennewick Man and the first Americans.
_ _
568
The story of Kennewick Man, introduced in Chapter 4, involves one of the earliest known human skeletons in the Americas. Discovered accidentally along the Columbia River in Kennewick, Washington, in 1996, the bones have been dated to 7500 B.C. Shortly after the discovery, a local newspaper story about the find initiated a series of ethical dilemmas, legal actions, court decisions, and publications concerning the questions that arise from the discovery of prehistoric human remains in the United States. Five Native American tribes in the area claimed these remains for reburial under the NAGPRA legislation. The Army Corps of Engineers took possession of the skeleton, which had been found on Corps lands, and agreed to return it to the tribes. Shortly thereafter, eight archaeologists and biological anthropologists filed an injunction to prevent the immediate return of the skeleton and to allow scientific investigation of this important find. The scientists argued that the remains were important and that they were not demonstrably Native American, questioning the legality of the tribes’ claim. This issue was in the courts and the media for almost a decade (Figure 12.5). In 2002, a judge ruled that the “remains were so old, and information as to his era so limited, that it is impossible to say whether the Kennewick Man is related to the present-day Tribal Claimants.” The judge also ordered that the remains be given over for scientific investigation. The tribes appealed this ruling, but an appeals court in 2004 upheld the judge’s decision. The issue now seems to have been settled. The ongoing conflict between science and belief is highlighted in such issues. The public battle over burial rights and the right to pursue knowledge about the past has both underscored and brushed aside the importance of the Kennewick find. This skeleton is important because of its antiquity and also because of the unusual nature of the remains. Questions raised by certain features of the skull demand further investigation to better understand the original inhabitants of the Americas. At the same time, the importance of learning more about the past of all of us is lost in the debate about who should control these remains and how they should be handled. Kennewick is a special case, a very old skeleton with distinct and unusual characteristics. Thousands of human remains have been excavated in the United States that are not as old and not as unusual. These remains are by law to be given to identifiable descendants who request their return. The issue of responsibility for human remains is one of the thorniest in archaeology. On the one hand, these materials are the remains of ancestors and significant, even sacred, in many cultures. On the other hand, the bones and tissues of these ancestors hold an
Chapter 12 The Past as Present and Future
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 569
enormous amount of information about the individual represented and about our common human past. As the Kennewick case shows, the issue is not easily resolved when diverse interests fight for control of such remains.
ETHICS IN ARCHAEOLOGY Ethics represent a common understanding among people about what is right and dutiful behavior in human society. Ethics are an important and growing concern in modern life. Ethical behavior in archaeology is increasingly complex, related to issues of heritage, native peoples, treatment of the dead, and the preservation of the past. Several books on the subject have appeared in recent years, and panels on ethical issues are frequently found at scientific conferences. In addition, national organizations have added ethical statements to their bylaws. The major professional organization for archaeologists in the United States is the Society for American Archaeology (SAA). Founded in 1934, the society has more than 6500 professional archaeologists and students as members today. The SAA seeks to stimulate interest and research in American archaeology, advocate and aid in the conservation of archaeological resources, encourage public access to and appreciation of archaeology, oppose all looting of sites and the purchase and sale of looted archaeological materials, and serve as a bond among those interested in the archaeology of the Americas. The promotion of ethical behavior is an important aspect of the society’s activities. The major principles of ethical practice have been set down by the SAA and are summarized in Table 12.4. These principles encompass the major concerns of archaeology. Stewardship concerns the protection of our common archaeological and cultural heritage; unless this heritage is protected, it will disappear. This principle of protection and preservation also extends to the collections and records that archaeologists accumulate. Accountability concerns archaeology’s interaction with the individuals or groups involved with particular archaeological sites or materials. These individuals and groups include the landowners of sites where archaeological projects may take place and Native American groups whose beliefs incorporate ancient artifacts and human remains.
Figure 12.6 Diver bringing a bucket from the surface to collect glass remains found on the Glass Wreck, an eleventh-century ship that sank off the coast of Turkey with over 200 pieces of glasswork.
THE RESPONSIBLE ARCHAEOLOGIST It is important to remember that archaeology is not a do-it-yourself activity. Amateur archaeologists in most cases work hand in hand with professionals. Most professional archaeologists have advanced degrees and have spent a great deal of time in the field and the classroom learning the many facets and skills of the discipline (Figure 12.6). Historical archives may be studied over and over again, but archaeological sites are nonrenewable resources. Excavations involve moving earth and all its contents from a site. Every excavation means the destruction of all or part of an _ The Responsible Archaeologist
569
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 570
TABLE 12.4 Eight Ethical Principles of the Society for American Archaeology 1. Stewardship. The archaeological record is irreplaceable, and it is the responsibility of all archaeologists to practice and promote stewardship of the archaeological record. 2. Accountability. Responsible archaeological research requires a commitment to consult with affected group(s) to establish a working relationship that can be beneficial to all parties involved. 3. Commercialization. The buying and selling of objects contributes to destruction of the archaeological record on the American continents and around the world. Archaeologists should discourage and avoid activities that enhance the commercial value of archaeological objects. 4. Public education and outreach. Archaeologists should work with the public to improve the preservation, protection, and interpretation of the record. 5. Intellectual property. A researcher may have primary access to original materials and documents for a limited and reasonable time, after which those materials and documents must be made available to others. 6. Public reporting and publication. The knowledge that archaeologists obtain in their investigations should be presented to the public. 7. Records and preservation. Archaeologists should work actively for the preservation of archaeological collections, records, and reports. 8. Training and resources. Archaeologists must ensure that they have the training, experience, facilities, and other support necessary to conduct a program of research.
archaeological site. It is for this reason, and in recognition that new methods and techniques in the future will be available to reveal more about the past, that archaeologists usually leave a substantial portion of a site undisturbed. All that is left when an excavation is over are the finds themselves, the unexcavated parts of the site, and the samples, photographs, drawings, measurements and other notes that the archaeologists made. Accurate notes and records of the layers, structures, and finds at a site are essential, not only for the investigator but also as a permanent archive of information that is available to others. In addition to keeping accurate records for a permanent archive, it is essential that archaeologists make their work public—known to other archaeologists and to the general public. Publication—making things public—in print or other media is part of the responsibility that archaeologists have to make their work and conclusions known so that the information can be checked and added to the general body of knowledge about the past. Books, articles in scientific journals, and reports are part of the permanent record that archaeology creates to document what is learned. In addition to scientific publication, archaeologists try to make their discoveries known to the general public through magazines, newspapers, television, and the Web. Hardly a day goes by without mention of a new discovery or revelation about the past in the popular media. Keeping the public informed is part of the responsibility of being an archaeologist. One of the popular means of keeping the public informed has been the creation of an Archaeology Week in many states (Figure 12.7). This is an annual event at which lectures, exhibitions, and other public activities provide more information to the public and raise awareness of the importance of archaeology. A third responsibility regards the commercialization of the past, and particularly the illegal trade in antiquities that remains a major force in the destruction of the past by looters and thieves. The archaeologist must discourage and avoid such activities. For example, an archaeologist should not provide estimates of the value of antiquities to collectors and retailers. Public education is an important role for archaeologists, intended to increase both interest and support. The more people who understand and appre-
_ _
570
Chapter 12 The Past as Present and Future
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 571
Figure 12.7 The 1992 poster for Archaeology Week in Arizona, one of the first states to initiate this program.
ciate the importance of the past, the better the job archaeology is doing. In addition, much of the financial support for research archaeology comes from government sources and ultimately from the public. The ethical principles of the SAA also concern intellectual property—researchers are entitled to the copyright on original ideas and information that they produce. At the same time, it is essential that these materials become available to other scholars and the public through publication in various media. Archaeological finds—artifacts, photographs, reports, and catalogs—should be stored under good conditions in perpetuity so that interested parties can study the evidence that has been recovered. Archaeological materials are normally stored in museums, university collections, and government facilities to provide a lasting record. The last ethical principle on the SAA list encourages archaeologists to have the training, resources, and support to be able to do their work at a high level. Museums and monuments are, of course, another way in which the public has access to the past. Archaeological materials are displayed in exhibitions in museums around the world to document the activities and achievements of ancestors and former inhabitants of the area. Living museums like Williamsburg, Virginia, invite the visitor to partake of the past, experiencing the architecture and activities of early America. National monuments in the United States and The Responsible Archaeologist
_
571
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 572
elsewhere protect the remains of past societies in the form of living places, architecture, mounds, and other structures that mark places of archaeological and historical interest (Figure 12.8). Museums and monuments serve both to inform the public of the past around them and to inspire more interest in the preservation and recovery of artifacts from former times.
CAREERS IN ARCHAEOLOGY w w w. m h h e . c o m / p r i c e i p 6 e
For a Web-based activity on the value of careers in archaeology, see the Internet exercises on your online learning center.
© 2004 by Sidney Harris. Reprinted with permission of ScienceCartoonsPlus.com.
_ _
572
There are many ways to participate in archaeology: watch a TV program, read a book, visit a museum, or volunteer at an archaeological site. For those with a more serious interest and a desire to pursue a career in archaeology, an advanced degree is normally required, either a master’s or, more likely, a PhD. Degrees in archaeology are offered by academic departments. There are many different kinds of archaeology, and their academic homes vary. In the United States, archaeology is usually part of the anthropology department, which combines archaeology with biological anthropology and cultural anthropology, all focused on the study of humans and culture. Archaeology may also be taught in a department of classics or art history or religious studies. In most of the rest of the world, archaeology has its own department or institute. Undergraduate students interested in archaeology usually take courses in anthropology, geography, geology, zoology, botany, soils, and the like to learn basic information about the discipline and related fields. Participation in fieldwork is an important activity at all levels of education in archaeology. Problems and questions of field research are complex, and gaining experience in a variety of situations and time periods is a great way to learn about the past. Inexperienced students often enroll in field schools—training programs in archaeological survey and excavation that are offered as credit courses by many universities in the summer. Fieldwork is often the crucible whereby students learn if archaeology is really the career they want to follow. Most undergraduate students apply to graduate school during their senior year or soon after graduation. In addition, there are older returning students who have decided they want to pursue an education in archaeology. Application to graduate schools usually requires an undergraduate transcript, letters of recommendation, scores from the Graduate Record Examination (GREs), and a statement of interest, along with the processing fee. Most graduate students start their studies with some specific interest in a time or place or subject matter. Most have an undergraduate background in archaeology and often some field experience. Advanced degrees require a major commitment in time, energy, and money. A master’s degree involves 1–3 years of coursework and may require a written thesis as well. Most programs include a comprehensive examination at the end of the master’s curriculum. The goal of a master’s program is to instill a broad general knowledge, both practical and theoretical, about the field of archaeology that will allow the student to perform well in the discipline. There are also more focused MA programs that teach specific skills, such as cultural resource management, human osteology, and geographic information systems (GIS). The PhD degree takes an average of 7.5 years, including the master’s, and is intended to provide specialized knowledge of particular areas of archaeology and to demonstrate exceptional research skills. Coursework is usually part of the curriculum, but the major part of PhD studies involves research. A PhD dissertation, the culmination of this degree program, is essentially a book describing the questions, methods, and results of a major research project. Archaeology is a fascinating career, in large part because the subject matter is highly diverse and highly human. There are so many times and places involved, and so many questions to be asked. Archaeology as a career accommodates an extraordinarily wide range of interests: archaeological chemistry, archaeozoology,
Chapter 12 The Past as Present and Future
pri31057_ch12.pp560-575.qxd
10/1/09
8:30 AM
Page 573
Figure 12.8 The ruins of Skara Brae, Orkney Islands, United Kingdom, a Neolithic settlement at the edge of northern Europe, 3000 B.C.
biblical archaeology, bioarchaeology, ceramic archaeology, classical archaeology, computer archaeology, ethnoarchaeology, experimental archaeology, geoarchaeology, historical archaeology, lithic archaeology, museum archaeology, paleoanthropology, paleoethnobotany, theoretical archaeology, and underwater archaeology. For those with a master’s degree or a PhD in hand, there are many directions to go in, although not a huge number of jobs. Archaeologists work in several kinds of places: as professors teaching and doing research in universities and colleges, as curators cataloging and researching artifacts in museums, as professional archaeologists in commercial firms doing fieldwork and writing reports on the impact of development on cultural resources (cultural resource management), and as government employees—rangers, researchers, and caretakers of the archaeology of federal properties. Cultural resource management (CRM) is a result of federal and state legislation in the United States over the past 50 years. In the 1960s and 1970s, the U.S. Environmental Protection Agency established requirements for an environmental impact study to determine whether important archaeological or historical sites were in danger of destruction prior to the start of federally funded construction. Both federal agencies and private corporations must provide information on the effect their construction projects may have on the history and prehistory of an area. This legislation has led to substantial growth in archaeology in both government and private sectors, with the creation of state and federal agencies and private companies to conduct and evalute these impact studies. Various kinds of construction—reservoirs, highways, sewage systems, and power lines, to name just a few—require such archaeological surveys and environmental impact statements. This kind of work is an important part of the field of archaeology. More than half of the archaeologists in the United States today are in CRM-related fields for private businesses, government agencies, and universities. These CRM projects usually involve three stages. An initial archive and field survey (Phase I) is made to see if archaeological or historical artifacts are present. If such materials are found, a second stage (Phase II) of more detailed testing and evaluation is undertaken. If these results are positive, a third stage (Phase III) may involve a full program of excavation and recovery or mitigation. Mitigation may involve moving the place of construction, rerouting rights-of-way, or doing more complete excavation. CRM projects vary greatly in size and scope. Many of these evaluations are relatively small scale and require only a few days or weeks for the determination of the impact of construction. Major building projects (e.g., dams, pipelines, and highways), on the other hand, may require years of fieldwork, analysis, and report writing.
Archaeology is one of those special fields where the public appeal and interest of the subject is combined with the intellectual pursuits of science: This is the time to extend the physical sciences and engineering into problem domains closer to the central interests of the non-scientific, to the light they increasingly throw on human nature and destiny. Archaeology is surely a major example of that linkage, along with astronomy and cosmology, so close to the old existential questions of all thinking humans. —Philip Morrison, MIT
cultural resource management The
survey and/or excavation of archaeological and historical remains threatened by construction and development.
Careers in Archaeology
573
_
pri31057_ch12.pp560-575.qxd
< ( : < V : y < p : W
10/1/09
8:30 AM
Images and Ideas The End
P
By now the accelerative thrust triggered by man has become the key to the entire evolutionary process on our planet. The rate and direction of the evolution of other species, their very survival, depends on decisions made by man. Yet there is nothing inherent in the evolutionary process to guarantee man’s own survival. Throughout the past, as successive stages of social evolution unfolded, man’s awareness followed rather than preceded the event. Today unconscious adaptation is no longer adequate. Faced with the power to alter the gene, to create new species, to populate the planets, or depopulate the earth, man must now assume conscious control of evolution itself. Avoiding future shock as he rides the waves of change, he must master evolution, shaping tomorrow to human need. Instead of rising in revolt against it, he must, from this historic moment on, anticipate and design the future. —Alvin Toffler (1970)
In my beginning is my end. —T. S. Eliot
w w w. m h h e . c o m / p r i c e i p 6 e
For more review material and study questions, see the self-quizzes on your online learning center.
_ _
574
Page 574
erhaps the best way to conclude this book is to ask, What have we learned? We know that archaeology includes stunning museum pieces, travel, and exotic locales. Archaeology is about places and things, method and theory, holes in the earth, open air, and dark storerooms in dusty museums. Archaeology is puzzles, mysteries, and detective work. Archaeology is intriguing methods for getting answers about the past and a stimulating career in which any number of interests can be explored by looking at the variety of times, places, and ways people lived in the past. Archaeology is good, dirty fun. Archaeology is also about diversity. It is very clear from the times and places in the past that we as humans have created many ways of solving the problems of survival and of improving the quality of life. The variety of ideas, technologies, and peoples that exist on the planet has ensured an increase in numbers, geographic expansion, and our success as a species. Archaeology allows us to see the variety of human societies and gain an appreciation of how humans are both different and alike in time and space. Archaeology is also about the oneness of humanity. Perhaps more than any other discipline, archaeology tells us that we are all members of the same family, traveling together on a miraculous journey through time. We take that journey with the plants and animals with which we share the planet. We all began in those primordial seas as replicating molecules. As Carl Sagan put it, we are all composed of that original starstuff that spewed across the new universe to form stars and planets and life. Another important lesson of archaeology has to do with the tempo of change. The rate of change in our societies and in ourselves is ever increasing. If we consider the passage of time as a straight line, the major milestones in the evolution of life occur more and more frequently as we approach the present. The complexity and rate of change in human prehistory are clearly accelerating. We took our first steps some 7–6 m.y.a., made our first tools 3–2 m.y.a., first buried our dead perhaps 100,000 years ago, painted on the walls of caves about 25,000 years ago, domesticated plants and animals 10,000 years ago, built cities and began to write on clay around 5000 years ago, and began to harness fossil fuels and machinery less than 500 years ago. The pace of life, science, and living has become so rapid that today almost everything around us changes in the short period between birth and death. It becomes more and more difficult to understand or cope with changes in the society that fostered us. We must deal with what philosopher and visionary Alvin Toffler has called future shock, dealing with the very rapid pace that will characterize the future. Perhaps the final lesson of archaeology is a lingering sense of hope about the future. Our strongest feelings from what we as archaeologists have learned remain a basic optimism for our species. In every way, we are artifacts, manufactured over a very long period of time, created by the experiences of our ancestors. We have been on the planet for several million years. In that time, we have evolved from a chimpanzee-like ape to man on the moon. We have expanded geographically and survived under a wide range of difficult conditions. There is an unusual quality about the human species—the enormous potential in the human intellect, with its remarkable inventiveness. A large brain and creativity managed to get us through a very long and difficult journey in the past. For that very reason, the future should be just as exciting.
Chapter 12 The Past as Present and Future
pri31057_ch12.pp560-575.qxd
p
10/1/09
8:30 AM
Page 575
DISCUSSION QUESTIONS 1. Why is it important to study the past? That is, why is archaeology important? 2. How can we protect our heritage? Why should we want to do so? 3. What are some key issues in the “ownership” of the past? What role does the government play here? 4. What are the ethical responsibilities associated with archaeology?
z
SUGGESTED READINGS For Internet links related to this chapter, please visit our Web site at www.mhhe.com/priceip6e.
Atwood, R. 2004. Stealing history: Tomb raiders, smugglers, and the looting of the ancient world. New York: St. Martin’s Press. Bahn, P. 1999. Bluff your way in archaeology. London: Oval Books. A lighthearted look at some of the myths and the truths of archaeology. Jameson, J. H. 1997. Presenting archaeology to the public: Digging for truths. Lanham, MD: Rowman & Littlefield. Strategies educators and archaeologists can use to present archaeological information to the public. Little, B. J. 2002. Public benefits of archaeology. Gainesville: University Press of Florida. Merriman, N., and T. Schadla-Hall. 2002. Public archaeology. London: Routledge.
Scarre, C., and G. Scarre. 2006. The ethics of archaeology. Cambridge: Cambridge University Press. Stille, A. 2003. The future of the past. London: Picador Books. A detailed expression of concern for the ongoing destruction of the past. Vitelli, K. D., and C. Colwell-Chanthphonh. 2006. Archaeological ethics, 2d ed. Blue Ridge Summit, PA: Altamira Press. An excellent summary of the issues that evoke ethical responses in archaeology. Zimmerman, L. J., K. D. Vitelli, and J. Hollowell-Zimmer, eds. 2003. Ethical issues in archaeology. Walnut Creek, CA: Altamira Press. A new consideration of major ethical issues in American archaeology.
_ Suggested Readings
575
pri31057_app.pp576.qxd
10/2/09
8:24 AM
Page 576
Appendix < Common Measurement Conversions and Equivalents
LENGTH OR HEIGHT
VOLUME
1 centimeter (cm) = 0.394 inch (in) 1 inch = 2.54 centimeters 1 meter (m) = 3.281 feet (ft) 1 foot = 0.305 meter 1 meter = 1.0936 yards (yd) 1 yard = 0.9144 meter 1 kilometer (km) = 0.6214 mile (mi) 1 mile = 1.6094 kilometers
1 cubic centimeter (cc) = 0.061 cubic inch (cu in) 1 cubic inch = 16.39 cubic centimeters 1 cubic meter (cu m) = 35.314 cubic feet (cu ft) 1 cubic foot = 0.0283 cubic meter 1 cubic meter = 1.308 cubic yards (cu yd)
APPROXIMATE EQUIVALENTS Average person Height of basketball basket High-diving platform Bowling alley Ten-story building Arc de Triomphe High ski jump Football field Washington Monument Golf course
Feet
Meters
5.5 10
1.7 3.0
32.8 60 100 164 196.8 300 555 20,000
10 18 30 50 60 91 169 6098
1 square centimeter (sq cm) = 0.155 square inch (sq in) 1 square inch = 6.452 square centimeters 1 square meter (sq m) = 10.764 square feet (sq ft) 1 square foot = 0.0929 square meter 1 square meter = 1.196 square yards (sq yd) 1 square yard = 0.8361 square meter 1 square kilometer (sq km) = 0.386 square mile (sq mi) 1 square mile = 2.59 square kilometers 1 hectare (ha) = 10,000 square meters 1 hectare = 2.47 acres 1 acre = 0.405 hectare
576
Appendix
Cubic Meters
Refrigerator Home bathroom UPS delivery truck School bus Large room Medium-size house Small church Modern oil tanker
2 10 25 50 100 1000 10,000 100,000
DENSITY
AREA
_ _
APPROXIMATE EQUIVALENTS
25 people/sq km 500 people/sq km 5000 people/sq km
= 0.2 person/soccer field = 65 people/sq mi = 4 people/soccer field = 1300 people/sq mi = 40 people/soccer field = 13,000 people/sq mi
APPROXIMATE EQUIVALENTS Average bedroom Soccer goal Doubles tennis court Basketball court Average house lot Baseball infield Olympic pool Football field Hockey rink Soccer field Union Square, New York City Average city block
Churchill Downs (horse racetrack) Indianapolis racetrack Central Park, New York City Manhattan Island
Length
Width
Square Feet
Square Meters
Acres
Hectares
12 ft 24 ft 36 ft 84 ft 80 ft 90 ft 50 m 300 ft 200 ft 100 m
10 ft 8 ft 78 ft 45.75 ft 80 ft 90 ft 21 m 60 ft 100 ft 80 m
120 192 2808 3843 6400 8100 11,298 18,000 20,000 86,080
11 18 261 357 595 753 1050 1673 1859 8000
0.0027 0.0045 0.0645 0.0882 0.1469 0.1859 0.2594 0.4132 0.4591 1.9760
0.0011 0.0018 0.0261 0.0357 0.0595 0.0753 0.1050 0.1673 0.1859 0.8000
650 ft 700 ft
650 ft 700 ft
422,500 490,000
39,250 45,539
9.6948 11.2481
3.9250 4.5539
Miles
Miles
Square Miles
Acres
Hectares
1.25 2.5
oval oval
0.018522 0.4112
11.849 263.057
4.797 106.501
799.663 14,074.060
323.750 5698.000
2.5 12.5
0.5 2.5
1.25 22
pri31057_glo.ppG1-G10.qxd
10/2/09
8:28 AM
Page G-1
<
GLOSSARY absolute dating A method of assigning archaeological dates in calendar years so that an age in actual number of years is known or can be estimated. accelerator mass spectrometry (AMS) A method of radiocarbon dating using an accelerator to count the individual isotopes of the carbon sample; advantages include small sample size, speed of counting, and accuracy. Acheulean A major archaeological culture of the Lower Paleolithic; named after the site of St. Acheul in France. A hallmark of the Acheulean is the handaxe. achieved status Social status and prestige attributed to an individual according to achievements or skills rather than inherited social position. See also ascribed status. acropolis (Greek, pl. acropoli) A raised complex of palaces and courtyards, especially in Mesoamerica and Greece. Adena A burial mound complex that developed in the Ohio River Valley toward the end of the last millennium B.C. adobe A mud mixture used to make sundried bricks for buildings in arid areas. adze A heavy, chisel-like tool. alpaca A domesticated South American herbivore with long, soft wool. altiplano (Spanish) The high-altitude plain between the eastern and western ridges of the Andes in Peru. Anasazi One of three major cultural traditions of the American Southwest during late prehistoric times. The Anasazi were centered in the northern Southwest, on the high plateau of the Four Corners region. Ancestral Pueblo One of the three major cultural traditions of the American Southwest during late prehistoric times. The Ancestral Pueblo were centered in the northern Southwest, on the high plateau of the Four Corners region. ancient DNA Genetic material preserved in archaeological remains of bones and plants that can be studied for information about past genetic relationships.
anthropogenic Describes an effect or process resulting from human activity. The creation of pasture from forest through intentional burning is an example. At times, human abandonment of a site or area (as well as initial occupation) also can set off environmental changes. anthropological archaeology (prehistory) Archaeological investigations that seek to answer fundamental questions about humans and human behavior. anthropomorphic Having human form or attributes. archaeoastronomy The study of ancient alignments and other aspects of the archaeological record and their relationship to ancient astronomical knowledge and events. archaeoethnobotany The study of plant remains from archaeological sites. archaeological record The body of material and information that survives for archaeologists to study. archaeology The study of the human past, combining the themes of time and change. archaeozoology The study of animal remains from archaeological sites. Archaic The term used for the early Holocene in the Americas, from approximately 6000 B.C. to 1500–1000 B.C. archipelago A group or chain of islands. artifact Any object or item created or modified by human action. ascribed status Social status and prestige attributed to an individual at birth, regardless of ability or accomplishments. See also achieved status. assemblage The set of artifacts and other remains found on an archaeological site or within a specific level of a site. association The relationship between items in an archaeological site. Items in association are found close together and/or in the same layer or deposit. It is often used for dating purposes, because items found in association are assumed to be of the same age. Atlantean column A carved human figure serving as a decorative or sup-
porting column, such as at the Mesoamerican site of Tula. atlatl A spearthrower, or wooden shaft, used to propel a spear or dart; first appeared in the Upper Paleolithic and also was used in the precontact Americas. australopithecine The generic term for the various species of the genus Australopithecus, including A. ramidus, A. afarensis, and A. africanus. bajo (Spanish) A broad, flat, clay-lined depression in the Maya Lowlands that fills with water during the rainy season. ball court An I-shaped or oval prehispanic structure, found throughout Mesoamerica and the southwestern United States, that was the site of ritual ballgames. Bandkeramik An archaeological culture of the early Neolithic in central Europe, referring to the style of pottery: linear bands of incised designs on hemispherical bowls. barrow An earthen mound covering a burial; found in prehistoric Europe and Asia. bier A stand on which a coffin or a corpse is placed. bifacial A flaked stone tool on which both faces or sides are retouched to make a thinner tool. binary-coded A system of information storage or processing in two states (such as 0 and 1). Computers process immense amounts of information through the manipulation of such sequences. biological anthropology The study of the biological nature of our nearest relatives and ourselves. bioturbation Activities of plants and animals in the earth, causing disturbance of archaeological materials. bipedalism The human method of locomotion, walking on two legs; one of the first human characteristics to distinguish the early hominins, as opposed to quadrupedalism, walking on four legs. blade A special kind of elongated flake with two parallel sides and a length at least twice the width of the piece.
Glossary
G-1
_
pri31057_glo.ppG1-G10.qxd
_ _
10/2/09
8:28 AM
The regular manufacture of blades characterized the Upper Paleolithic, with an efficient way of producing mass quantities of cutting edge. blank A piece, also known as a preform, that can be made into several different shapes or forms, usually in stone or metal. bonobo A small species of chimpanzee, closely related to humans. bow-drill A device for perforating beads or other small objects, in which a bow is used to rotate the shaft of the bit. brazier An open pan or ceramic vessel used for holding hot coals. breccia The accumulated materials from cave deposits that harden into a conglomerate rock, including sediments, rocks, and animal bones. bronze A mixture of copper and tin or arsenic to make a hard, durable metal. brow ridge That part of the skull above the eye orbits. This ridge of bone was particularly pronounced in the early hominins, when cranial capacity was less and the forehead absent or sloping. Brow ridges are largely absent in Homo sapiens sapiens. bulla (Latin, pl. bullae) A hollow clay sphere or envelope used to enclose clay tokens in ancient Mesopotamia. burin A stone tool with right-angle edges, used for planing and engraving. cacao A bean of the cacao tree, native to Mesoamerica; used to make chocolate. Cacao beans also were used as money by the Aztecs. cache A collection of artifacts, often buried or associated with constructed features, that has been deliberately stored for future use. calibrated dates Dates resulting from the process of calibration, the conversion of radiocarbon years to calendar years, by means of a curve or formula derived from the comparison of radiocarbon dates and tree rings from the bristlecone pine. Calibration extends approximately 6000 years into the past. camelid A ruminant mammal—such as camel, llama, and extinct related forms—having long legs and two toes. cannibalism The practice of humans eating human flesh. carnelian A red or reddish variety of chalcedony (a translucent variety of quartz) used in jewelry. cayman A tropical South American alligator. celt An implement shaped like a chisel or an axe; may be made of stone or metal.
G-2
Glossary
Page G-2
cenotaph A grave that does not contain a skeleton. cenote The Maya word for a sinkhole, a natural well in the Yucatán that provides water for drinking and bathing. chac mool (Maya) A life-size stone figure in a reclining position, with flexed legs and head raised and turned to one side. Chac mools served as altars and were often placed in temple doorways to receive offerings. charnel house A house in which the bodies of the dead are placed. chert A dull-colored, subtranslucent rock resembling flint that was often used for making flaked stone tools. chicha A South American beer made from maize. Chichimec A term loosely applied to the peoples who lived beyond the northern limits of Mesoamerica; nomadic people, considered to be uncivilized barbarians. chinampa (Spanish) An agricultural field created by swamp drainage or landfill operations along the edges of lakes. This intensive form of agriculture was especially prevalent in the Basin of Mexico but also was used elsewhere in the Central Highlands of Mexico. chlorite A kind of green stone that resembles mica. circumscription The process or act of being enclosed by either environmental boundaries, such as mountains, oceans, and rivers, or social boundaries, such as neighboring groups of people. citadel A hilltop fortress, the characteristic settlement of the ruling elite of Mycenaean civilization, 1700– 1100 B.C. Clactonian A term used for assemblages from the Lower Paleolithic, lacking handaxes and characterized by large flakes with heavy retouching and notches. class A relationship of inequality between members of society in which status is determined by membership in a level or class. The caste system in India is a prime example. cleaver A companion tool of the Acheulean handaxe. Cleavers have a broad leading edge, whereas handaxes come to a point. Clovis An archaeological culture during the Paleoindian period in North America, defined by a distinctive type of fluted point; named for the original find spot near Clovis, New Mexico.
coca A native Andean shrub whose dried leaves are chewed as stimulants. codex (Latin, pl. codices) A handpainted book on bark paper or animal skins folded like a screen. In Mesoamerica, codices, which record historical, religious, and tribute information, were made both before and after the Spanish conquest. comal (Spanish) A flat, ceramic griddle used for cooking tortillas. conquistador (Spanish) A conqueror; refers to the Spanish explorers who conquered Mexico in the early 1500s and also ventured into the southern United States. context The association and relationships between archaeological objects that are in the same place. coprolite Fossilized feces. cord-marking A decorative technique in Jomon Japan and elsewhere, in which cord or string is wrapped around a paddle and pressed against an unfired clay vessel, leaving the twisted mark of the cord. core The stone from which other pieces or flakes are removed. Core tools are shaped by the removal of flakes. cosmology The worldview of a group or society, encompassing their understanding of the universe, their origins and existence, and nature. craft specialists (or craft specialization) Individuals involved in part- or fulltime activities devoted to the production of a specific class of goods, often highly valued. Examples include metalsmiths, potters, and bead makers. cultigen A cultivated plant. cultivation The human manipulation or fostering of a plant species (often wild) to enhance or ensure production, involving such techniques as clearing fields, preparing soil, weeding, protecting plants from animals, and providing water to produce a crop. cultural anthropology The study of living peoples and the shared aspects of the human experience. cultural resource management The survey and/or excavation of archaeological and historical remains threatened by construction and development. culture A uniquely human means of nonbiological adaptation; a repertoire of learned behaviors for coping with the physical and social environments. cuneiform A writing system of ancient Mesopotamia using a series of wedge-
pri31057_glo.ppG1-G10.qxd
10/2/09
8:28 AM
Page G-3
shaped marks to convey a message or text. cutmark A trace left on bone by a stone or metal tool used in butchering a carcass; one of the primary forms of evidence for meat-eating by early hominins. Cyclopean A term describing the huge stone walls of Mycenaean tombs and fortresses; from Cyclops, the mythical giant. danzante (Spanish) Dancer; a life-size carving of a captive or a prisoner of war depicted in bas-relief on stone slabs at San José Mogote and Monte Albán, Oaxaca. débitage A term referring to all the pieces of shatter and flakes produced and not used when stone tools are made. dendrochronology The study of the annual growth rings of trees as a dating technique to build chronologies. division of labor The differentiation of economic roles in a specific sociopolitical context. dolmen A generic term for a megalithic tomb or chamber with a roof. domestication The taming of wild plants and animals by humans. Plants are farmed and become dependent on humans for propagation; animals are herded and often become dependent on their human caretakers for food and protection. dryopithecine The generic term for the Miocene fossil ancestor of both the living apes and modern humans, found in Africa, Asia, and Europe. ecofact Any of the remains of plants, animals, sediments, or other unmodified materials that result from human activity. economy The management and organization of the affairs of a group, a community, or an establishment to ensure their survival and productivity. edge hypothesis The theory that the need for more food was initially felt at the margins of the natural habitat of the ancestors of domesticated plants and animals; a revised version of the population pressure hypothesis about the origins of agriculture. effigy A representation or image of a person or an animal. egalitarian A term that refers to societies lacking clearly defined status differences between individuals, except for those due to sex, age, or skill. See also hierarchical.
E group An arrangement of buildings designed to mark the position of the rising sun during important solar events, such as equinoxes and solstices, in Mesoamerica. El Niño (Spanish) A warm-water countercurrent that periodically appears off the Peruvian coast, usually soon after Christmas, and alters the normal patterns of water temperature, flow, and salinity. These changes diminish the availability of nutrients to marine life, causing large schools of fish and flocks of seabirds to either migrate or die. emblem glyph A set of Maya hieroglyphs; generally, each emblem glyph is specific to a given Classic Maya city. Although most Maya epigraphers agree that emblem glyphs have a geographic referent, they do not agree on whether such glyphs stand for a place or for the royal family that ruled the place. empire A union of dispersed territories, colonies, states, and unrelated peoples under one sovereign rule. endocast A copy or cast of the inside of a skull, reflecting the general shape and arrangement of the brain and its various parts. epigraphy The study of inscriptions. epiphysis The end of a long bone in humans and other mammals that hardens and attaches to the shaft of the bone with age. epoch A subdivision of geological time, millions of years long, representing units of eras. equinox A time when the sun crosses the plane of the equator, making the night and the day the same length all over the earth, occurring about March 21 and September 22. era A major division of geological time, tens or hundreds of millions of years long, usually distinguished by significant changes in the plant and animal kingdoms; also used to denote later archaeological periods, such as the prehistoric era. estrus The cycle of female sexual receptivity in many species of animals. estuary A low area along a coast where the wide mouth of a river meets the sea and the waters of the two mix. ethnocentrism Evaluating other groups or societies by standards that are relevant to the observer’s culture. ethnography The study of human cultures through firsthand observation.
ethnohistory The study of ancient (often non-Western) cultures using evidence from documentary sources and oral traditions, and often supplemented with archaeological data. Traditionally, ethnohistorians have been concerned with the early history of the Americas, the time of European contact, and later settlement and colonization by Europeans. evolution The process of change over time resulting from shifting conditions of the physical and cultural environments, involving mechanisms of mutation and natural selection. Human biology and culture evolved during the Late Miocene, Pliocene, Pleistocene, and Holocene. excavation The exposure and recording of buried materials from the past. extrasomatic Literally, “outside the body”; nonbiological, nongenetic. facade The face, or front, of a building. feature An immovable structure or layer, pit, or post in the ground having archaeological significance. Fertile Crescent An upland zone in Southwest Asia that runs from the Levant to the Zagros Mountains, with adequate rainfall and many wild species that were domesticated. fieldwork The search for archaeological sites in the landscape through surveys and excavations. flake A type of stone artifact produced by removing a piece from a core through chipping. flint A fine-grained, crystalline stone that fractures in a regular pattern, producing sharp-edged flakes; highly prized and extensively used for making flaked stone tools. flintknapping The process of making chipped stone artifacts; the striking of stone with a hard or soft hammer. floodwater farming A method of farming that recovers floodwater and diverts it to selected fields to supplement the water supply. flotation A technique for the recovery of plant remains from archaeological sites. Sediments or pit contents are poured into water or heavy liquid; the lighter, carbonized plant remains float to the top for recovery, while the heavier sediments and other materials fall to the bottom. fluted point The characteristic artifact of the Paleoindian period in North America. Several varieties of fluted points were used for hunting large Glossary
G-3
_
pri31057_glo.ppG1-G10.qxd
_ _
10/2/09
8:28 AM
game. The flute refers to a large channel flake removed from both sides of the base of the point to facilitate hafting. Folsom An archaeological culture during the Paleoindian period in North America, defined by a distinctive type of fluted point and found primarily in the Great Plains. fossil The mineralized bone of an extinct animal. Most bones associated with humans in the Pliocene, Pleistocene, and Holocene are too young to have been mineralized, but the term fossil skull or fossil bone is often used generically in those cases as well. frieze A decorative band or feature, commonly ornamented with sculpture, usually near the top of a wall. galena A common heavy mineral that is the principal ore of lead. gallery grave A megalithic tomb lacking an entrance passage; the burial room or rooms form the entire internal structure. Gallery graves are found in Neolithic western Europe. gazelle One of several species of smallto-medium, swift, and graceful antelopes native to Asia and Africa. geoglyph Ground markings, such as the lines and life-form representations found in the Nazca desert. geomorphic Having the form or attributes of surface features of the earth or other celestial bodies. glacial A cold episode of the Pleistocene, in contrast to a warmer interglacial period; also called an ice age. The classic European sequence of the Günz, Mindel, Riss, and Würm glacials has recently been revised, with the recognition of a large number of cold/warm oscillations in the Pleistocene. glaciation The expansion of continental glacial ice during a period of cold climate. glume The tough seed cover of many cereal kernels. In the process of the domestication of wheat, the tough glume became more brittle, making threshing easier. glyph (Greek) A carving; a drawn symbol in a writing system that may stand for a syllable, a sound, an idea, a word, or a combination of these. See also emblem glyph. gorget A circular ornament, flat or convex on one side and concave on the other, usually worn over the chest.
G-4
Glossary
Page G-4
grave goods The items that are placed in graves to accompany the deceased. ground-penetrating radar (GPR or georadar) An instrument for remote sensing or prospecting for buried structures using radar maps of subsoil features. guano Bird excrement. half-life A measure of the rate of decay in radioactive materials; half the radioactive material will disappear within the period of one half-life. hammerstone A stone used to knock flakes from cores. handaxe A large, teardrop-shaped stone tool bifacially flaked to a point at one end and a broader base at the other. The characteristic artifact of the Lower Paleolithic; for generalpurpose use that continued into the Middle Paleolithic. handedness Preferential use of the right or the left hand; related to the organization of the brain in two hemispheres. hard-hammer technique A flintknapping technique for making stone tools by striking one stone, or core, with another stone, or hammer. See also soft-hammer technique. hematite A common heavy mineral that is the principal ore of iron. hemp A tall annual plant whose tough fibers are used to make coarse fabrics and ropes. henge A monument defined by the presence of an enclosure, usually made by a circular ditch and bank system, up to 500 m in diameter. Henges were erected during the Neolithic and early Bronze Age in western Europe. hierarchical A term referring to societies that have a graded order of inequality in ranks, statuses, or decision makers. See also egalitarian. hieroglyph Originally, the pictographic script of ancient Egypt; any depictive, art-related system of writing, such as that of Mesoamerica; also may refer to an individual symbol. historical archaeology Archaeology in combination with the written record. Hohokam One of three major cultural traditions of the American Southwest during late prehistoric times. The Hohokam were centered in the deserts of southern Arizona. hominid An obsolete term that refers to the human members of the primates, both fossil and modern forms.
hominin A term that refers to the human, chimp, and gorilla members of the primates, both fossil and modern forms. hominoid A descriptive term for any human or ape, past or present, characterized by teeth shape, the absence of a tail, and swinging arms. Hominoidea The taxonomic group (family) that includes the human and ape members of the primates, both fossil and modern forms. Hopewell Interaction Sphere A complex network involving the exchange of goods and information that connected distinct local populations in the midwestern United States from approximately 100 B.C. to A.D. 400. horizon A widely distributed set of cultural traits and artifact assemblages whose distribution and chronology suggest they spread rapidly. A horizon is often composed of artifacts associated with a shared symbolic or ritual system. household archaeology The archaeological analysis of past houses and associated residential remains to learn about domestic life and activities. huaca (Quechua) An Andean word for pyramid. hunter-gatherer A hunter of large wild animals and gatherer of wild plants, seafood, and small animals, as opposed to farmers and food producers. Hunting and gathering characterized the human subsistence pattern before the domestication of plants and animals and the spread of agriculture. Hunter-gatherers are also known as foragers. hyoid bone A delicate bone in the neck that anchors the tongue muscles in the throat. iconography The study of artistic representations or icons that usually have religious or ceremonial significance. ideograph A written symbol that represents an abstract idea rather than the sound of a word. See also pictograph. ideology A conceptual framework by which people structure their ideas about the order of the universe, their place in that universe, and their relationships among themselves and with objects and other forms of life around them. incensario (Spanish) An incense burner made of pottery and sometimes stone, used in Mesoamerican religious and political ceremonies.
pri31057_glo.ppG1-G10.qxd
10/15/09
11:07 AM
inflorescence The flowering part of a plant. intaglio An engraving in stone or other hard material that is depressed below the surface; an impression of the design produces an image in relief. interglacial A warm period of the Pleistocene, in contrast to a colder period called a glacial. isotope One of several atomic states of an element; for example, carbon occurs as 12C, 13C, and 14C, also known as carbon-14 or radiocarbon. isotopic technique A method for absolute dating that relies on known rates of decay in radioactive isotopes, especially carbon, potassium, and uranium. jasper A high-quality flint, often highly colored; often used as a raw material for the manufacture of stone tools, beads, and other ornaments. jet A compact, black coal that can be highly polished; used to make beads, jewelry, and other decorative objects. Jomon The archaeological culture of late Pleistocene and early Holocene Japan; primarily associated with groups of hunter-gatherers, but recent evidence suggests that these groups were practicing some rice cultivation. jujube A small, edible fruit from an Asian tree of the buckthorn family. The fruit has one seed in the center, somewhat like a cherry. kiln A furnace or oven for baking or drying objects, especially for firing pottery. kiva A semisubterranean ceremonial room found at sites throughout the American Southwest. krater A large metal vessel for mixing and storing wine, traded over a large part of Europe during the Iron Age. lactational amenorrhea The suppression of ovulation and menstruation during breast-feeding. laguna (Spanish) Lagoon; a man-made depression in Mesoamerica that may have begun as a borrow pit for the construction of an earthen mound. Lagunas were often lined with waterproof bentonite blocks and may have been used for ritual bathing. lake dwelling Former name for Early Neolithic lakeshore settlements originally thought to have been built over the water. lapidary Of or related to the practice of working or cutting precious or semiprecious stone.
Page G-5
lapis lazuli A semiprecious stone of deep blue; used and traded widely in antiquity in the form of beads, pendants, and inlay. lateralization The division of the human brain into two halves. One side controls language; the other regulates perception and motor skills. leguminous plants Vegetables used as food. Levallois A technique for manufacturing large, thin flakes or points from a carefully prepared core; first used during the Lower Paleolithic and remaining common during the Middle Paleolithic. The method wasted flint and was generally not used in areas of scarce raw materials. Levant A mountainous region paralleling the eastern shore of the Mediterranean, including parts of the countries of Turkey, Syria, Lebanon, and Israel. lintel A horizontal beam of wood or stone that supports the wall above a doorway or window. lithic Pertaining to stone or rock. llama A woolly South American camelid; used as a beast of burden. locomotion A method of animal movement, such as bipedalism. loess Wind-blown silt deposited in deep layers in certain parts of the Northern Hemisphere. lomas (Spanish) Vegetation that is supported by fog in otherwise arid environments. Long Count The Classic Maya system of dating that records the total number of days elapsed from an initial date in the distant past (3114 B.C.). The system is based on multiples of 20 beginning with the kin (1 day), uinal (20 kins or 20 days), tun (18 uinals or 360 days), katun (20 tuns or 7200 days), and baktun (20 katuns or 144,000 days). longhouse A wooden structure that is considerably longer than it is wide that served as a communal dwelling, especially among native North Americans in the Northeast and on the Northwest Coast. lost wax casting A technique for casting metal in which a sand or clay casing is formed around a wax sculpture; molten metal is poured into the casing, melting the wax. The cooling metal takes on the shape of the “lost” wax sculpture preserved on the casing.
macaw Any of several varieties of parrots from Mexico and Central and South America that were prized for their colorful feathers. magnetite A black iron oxide that can be polished to a lustrous surface. maguey Any of several species of aridenvironment plants with fleshy leaves that conserve moisture. The fiber and needles of magueys were used to make rope and clothing in Mesoamerica and the southwestern United States. mano The hand-held part of a stonemilling assembly for grinding maize or other foods. marketing An exchange system that frequently involves currencies and generally extends beyond close kinsmen and a small group of trading partners. Market participants try to minimize their costs and maximize their returns to make a profit. Maya Blue A steadfast blue pigment made by fusing an extract from the plant indigo with a fine white clay, palygorskite. The Maya and other Mesoamerican peoples applied this pigment on a range of materials, including ceramics, sculptures, and murals. megalith A large stone monument. menhir A large, standing stone, found either alone or collectively in lines. Mesoamerica The region consisting of central and southern Mexico, Guatemala, Belize, El Salvador, and the western parts of Honduras and Nicaragua that was the focus of complex, hierarchical states at the time of Spanish contact. The people of this area shared a basic set of cultural conventions. Also called Middle America. Mesolithic The period of time of huntergatherers in Europe, North Africa, and parts of Asia between the end of the Pleistocene and the introduction of farming; the Middle Stone Age. Mesopotamia The flat plain between the Tigris and Euphrates rivers in southern Iraq where the world’s first civilization developed. mesquite A tree or shrub of the southwestern United States and Mexico whose beanlike pods are rich in sugar. metallurgy The art of separating metals from their ores. metate The stone basin, often troughshaped, or lower part of a stonemilling assembly for grinding maize or other foods.
_ Glossary
G-5
pri31057_glo.ppG1-G10.qxd
_ _
10/2/09
8:28 AM
mica A colored or transparent mineral silicate that readily separates into very thin sheets. Mica was carved to make ornaments and crushed as an inclusion to clay in the fabrication of pottery. microband A small family group of hunter-gatherers. midden An accumulated pile of trash and waste materials near a dwelling or in other areas of an archaeological site. Milankovitch forcing A term describing the phenomenon considered to be the prime reason for glacial fluctuations and climatic change. Changing factors are the distance between the earth and the sun and the tilt of the earth’s axis, which play major roles in the amount of sunlight reaching the earth, atmospheric temperature, and the expansion and retreat of continental glaciation. The cyclical nature of variation in these factors was recognized by Yugoslav mathematician Milutin Milankovitch. millennium A period of 1000 years. Mississippian The collective name applied to the agricultural societies that inhabited portions of the eastern United States from approximately A.D. 800–1700. Mississippian peoples constructed earthen platform mounds and shared certain basic cultural conventions. mit’a system A means of tribute in prehispanic Andean South America that involved the use of conscripted laborers to complete discrete organizational tasks. mitmaq A system of colonization used by the Inca to minimize provincial rebellion by moving people around to break up dissident groups. mitochondrial DNA Genetic material in the mitochondria of human cells that mutates at a relatively constant rate. Because mitochondrial DNA is inherited only from the mother, it provides an unaltered link to past generations. Mogollon One of three major cultural traditions of the American Southwest during late prehistoric times. The Mogollon were centered in the mountainous areas of southeastern Arizona and southwestern New Mexico. monochrome One color; describing pottery decorated with only one color that contrasts with the underlying color of the paste of the vessel. montaña (Spanish) Mountain, specifically referring to the wet, tropical slopes of the Amazonian Andes.
G-6
Glossary
Page G-6
mortar A bowl-shaped grinding tool, used with a wood or stone pestle for grinding various materials. motif A recurring thematic design element in an art style. Mousterian A term describing the stone tool assemblages of the Neanderthals during the Middle Paleolithic, named after the site of Le Moustier in France. See also Acheulean. multivallate A term describing complex defenses of multiple ditches and ramparts at large Iron Age hillforts. mural art One of the two major categories of Paleolithic art, along with portable art. Mural art consists of painting, engraving, and sculpting on the walls of the caves, shelters, and cliffs of southwestern Europe; one of the hallmarks of the Upper Paleolithic. m.y.a. Abbreviation for millions of years ago. natural habitat hypothesis The theory about the origins of agriculture associated with Robert Braidwood, suggesting that the earliest domesticates appeared in the area that their wild ancestors inhabited. necropolis (Greek) Cemetery. Neolithic The period of time of early farmers with domesticated plants and animals, polished stone tools, permanent villages, and often pottery; the New Stone Age. net-sinker A small weight attached to fishing nets. nome A geographic province incorporated within the ancient Egyptian state. oasis hypothesis The theory about the origins of agriculture associated with V. Gordon Childe and others, suggesting that domestication began as a symbiotic relationship between humans, plants, and animals at oases during the desiccation of Southwest Asia at the end of the Pleistocene. obsidian Translucent, gray to black or green, glasslike rock from molten sand; produces extremely sharp edges when fractured and was highly valued for making stone tools. oca A South American wood sorrel (Oxalis crenata) that is cultivated for its edible tuber. occipital bun A distinctive shelf or protrusion at the base of the skull; a feature usually associated with Neanderthals.
Oldowan The name given to the assemblages of early pebble tools and flakes belonging to the Basal Paleolithic, derived from Olduvai. Olmec The Aztec name for the late prehispanic inhabitants of the Gulf Coast region of Mexico. This term has been extended by archaeologists to describe the sites, monuments, and art found in the same region during the Formative period. Aspects of this art style and related motifs had a wider distribution across Mesoamerica during the Early and Middle Formative periods (1150–700 B.C.). This broader distribution is called the Olmec Horizon. oppidum A massive fortification in western Europe, often on a hilltop or a bluff, built for defensive purposes during the Iron Age; described in some detail and often conquered by the Romans. oracle bone An animal bone with cracks (due to heating) or other markings, used to foretell the future. organization The arrangements between individuals and groups in human society that structure relationships and activities. oxygen isotope ratio The ratio of different isotopes of oxygen in ocean water, varying with the temperature of the water; measured in seashells and used as an indicator of temperature change over time. paleoanthropology The branch of anthropology that combines archaeology and physical anthropology to study the biological and behavioral remains of the early hominins. Paleoindian The period of large-game hunters in North America at the end of the Pleistocene. Paleoindian remains are characterized by the presence of fluted points and frequently the bones of extinct animals. Paleolithic The first period of human prehistory, extending from the time of the first tools, more than 2.5 m.y.a., until the end of the Pleistocene, 10,000 years ago. Characterized by the use of flaked stone tools, it is also known as the Old Stone Age. palisade A fence of posts or stakes erected around a settlement for defensive purposes. panpipe A wind instrument consisting of bound sets of short pipes in graduated lengths. pantheon The officially recognized gods of a people.
pri31057_glo.ppG1-G10.qxd
10/2/09
8:28 AM
Page G-7
papyrus A tall marsh plant, or reed, of the Nile Valley that the ancient Egyptians cut into strips and pressed into a kind of paper to write on. Paranthropus Genus of early hominins, contemporary with Australopithecus, that includes boisei and robustus as species. passage grave A megalithic tomb entered via a long, low, narrow passage that opens into a wider room, generally near the center of the structure. pastoralist An animal herder. Pastoralism is a subsistence strategy generally associated with a mobile lifeway. patrilocal Describing a residence pattern in which married couples live with or near the husband’s family. pectoral A large ornament worn across the chest, especially for defensive purposes. percussion flaking A technique for producing stone artifacts by striking or knapping crystalline stone with a hard or soft hammer. See also pressure flaking. petroglyph A drawing that has been carved into rock. petty state A small, socially stratified political unit prevalent in Mesoamerica at the time of the Spanish conquest. Similar political formations have been found in other regions as well. phonetic Pertaining to the sounds of speech. pictograph A written or painted symbol that more or less portrays the represented object. See also ideograph. pipal tree A species of fig tree on the South Asian subcontinent that has had sacred significance for many cultures and religions throughout the region for thousands of years. pithos A large clay storage jar. pithouse A prehistoric semisubterranean dwelling in which the lower parts of the walls were the earthen sides of a shallow pit; the top parts of the walls often consisted of a framework of poles intertwined with small twigs, covered with mud. Plio-Pleistocene A term used to describe the time between the appearance of the earliest hominins during the Pliocene and the beginning of the Pleistocene. pochteca A privileged, hereditary guild of long-distance Aztec traders. polychrome Multicolored; describing pottery that has been decorated with three or more colors.
polygynous Having more than one mate. population pressure hypothesis Lewis Binford’s theory that population increase in Southwest Asia upset the balance between people and food, forcing people to turn to agriculture as a way to produce more food. porphyry An igneous rock with visible quartz or feldspar crystals embedded in a finer-grained base. portable art One of the two major categories of Paleolithic art, along with mural art. Portable art includes all decorated materials that can be moved or carried; found throughout Europe and much of Eurasia. post mold The circular remains, often just a dark stain in the soil, of a wooden post that formed part of the frame of a prehistoric structure; also called a posthole. potassium-argon dating See radiopotassium dating. potlatch A large feast among Northwest Coast Native Americans that included the display and dispersal of accumulated wealth to the assembled guests. potsherd A fragment of a clay vessel or object. prehistory In general, the human past; specifically, the time before the appearance of written records. pressure flaking A technique for producing stone artifacts by removing flakes from a stone core by pressing with a pointed implement. See also percussion flaking. primary context (in situ) An object found where it was originally located in antiquity, not redeposited. primate The order of animals that includes lemurs, tarsiers, monkeys, apes, and humans; characterized by grasping hands, flexible limbs, and a highly developed sense of vision. provenience The place of origin for archaeological materials, including location, association, and context. pueblo A stone-masonry complex of adjoining rooms found in the American Southwest. puna (Spanish) High grassland plateaus in the Peruvian Andes. quern A stone grinding surface for preparing grains and other plant foods and for grinding other materials. quinoa A pigweed (Chenopodium quinoa) of the high Andes. Seeds of the plant were ground and used as food in the past and still are today.
quipu The Inca word for an elaborate knotted-string device used by the Inca and other peoples in Peru for record keeping. A quipu consists of a horizontal cord from which a series of smaller knotted strings hang. The placement, color, and nature of the knots on the cords convey numbers and other information. rachis The stem that holds seeds to the stalk in wheat and other plants; changes from brittle to tough when wheat is domesticated. radiocarbon dating An absolute dating technique based on the principle of decay of the radioactive isotope of carbon, 14C; used to date archaeological materials within the past 40,000 years. radiopotassium dating An absolute dating technique based on the principle of decay of the radioactive isotope of potassium, 40K; used to date materials ranging in age from 500,000 years old to the age of the oldest rocks in the universe. Also called potassium-argon dating. ramón A tree that grows abundantly in the tropical forests of the Maya Lowlands and bears an edible fruit; also called breadnut. rank A relationship of inequality between members of society in which status is determined by kinship relations of birth order and lineage. reciprocity The exchange of goods between known participants, involving simple barter and face-to-face exchanges. redistribution The accumulation and dispersal of goods through a centralized agency, individual, or institution. red ochre An iron mineral that occurs in nature; used by prehistoric peoples in powdered form as a pigment for tanning animal skins; often found in burials from the late Paleolithic and Mesolithic. reduction technique In archaeology, a manufacturing process involving the removal (as opposed to the addition) of materials from a core that becomes the finished product; includes techniques such as flintknapping and wood carving. relative dating A technique used to estimate the antiquity of archaeological materials, generally based on association with materials of known age or simply to say that one item is younger or older than another. repoussé (French) The process of forming a raised design on a thin sheet of Glossary
G-7
_
pri31057_glo.ppG1-G10.qxd
_ _
10/2/09
8:28 AM
metal by placing it over a mold and hammering it in place. retouch The shaping or sharpening of stone artifacts through percussion or pressure flaking. rhizome An edible, rootlike subterranean plant stem. robust “Big-boned,” heavy, thick-walled skeletal tissue. Robust early hominins also had very large teeth. roof comb An architectural feature, frequently carved with glyphs and figures, that is placed on the top of Mesoamerican temples. sacbe The Maya word for a raised causeway constructed of stone blocks and paved with gravel and plaster. sarcophagus A stone coffin, usually decorated with sculpture and/or inscriptions. scapulimancy The ancient practice of seeking knowledge by reading cracks on bones. Symbols were written on an animal’s scapula (shoulder blade); the bone was heated until a series of cracks formed; then diviners interpreted the pattern of cracking to foretell the future. scepter A staff or baton borne by a ruler as an emblem of his or her position and authority. scheduling The process of arranging the extraction of resources according to their availability and the demands of competing subsistence activities. seal stamp A piece of inscribed stone used by administrators to impress a symbol on wet pieces of clay or bitumen in order to keep track of goods. seasonality The changing availability of resources according to the different seasons of the year. sedentism Living in permanent, yearround contexts, such as villages. serpentine A stone of dull green color that often has a mottled appearance. setaria A wild grass with edible seeds. sexual dimorphism A difference in size between the male and female members of a species; for example, male gorillas are significantly larger than females. sexual division of labor The cooperative relationship between the sexes in hunter-gatherer groups involving different male and female task activity. shaduf An Egyptian bucket-and-lever lifting device that enables one to raise water a few feet from a well or ditch onto fields and gardens.
G-8
Glossary
Page G-8
shaft grave A vertical tunnel cut into rock and holding the tombs of Mycenaean elite. shaman An anthropological term for a spiritualist, curer, or seer. shattering A natural mechanism of seed dispersal. shell midden A mound of shells accumulated from human collection, consumption, and disposal; a dump of shells from oysters, clams, mussels, or other species found along coasts and rivers, usually dating to the Holocene. shicra The Inca word for meshed bags containing rocks, used as fill in the construction of ancient Andean structures. sickle A tool for cutting the stalks of cereals, especially wheat. Prehistoric sickles were usually stone blades set in a wood or antler handle. sickle polish A clear polish that forms along the edges of flakes and blades that are used to cut reeds, grass, wheat, and other long-stemmed plants. site The accumulation of artifacts and/or ecofacts, representing a place where people lived or carried out certain activities. slash and burn A type of farming in which the ground is cleared by cutting and burning the vegetation on the spot. The burned vegetation serves as a natural fertilizer. The field is farmed until yields decrease; then it is allowed to lie fallow. Also called swidden farming. slate A fine-grained rock, with a dull, dark bluish-gray color, that tends to split along parallel cleavage planes, often producing thin plates or sheets. soapstone A soft stone with a soapy feel that is easy to carve; often referred to as steatite. social hypothesis The theory that domestication allowed certain individuals to accumulate food surplus and to transform those foods into more valued items, such as rare stones or metals, and even social alliances. sodality An alliance or association among some members of a society, often based on age and sex, with a specific function. Sodalities can be ceremonial, political, or economic; examples include dance societies, warrior groups, sororities, clubs, and fraternal organizations. soft-hammer technique A flintknapping technique that involves the use of a hammer of bone, antler, or wood,
rather than stone. See also hardhammer technique. solifluction A phenomenon in which freezing and thawing of the ground results in slippage of the surface. solstice The time of year when the sun is at its greatest distance from the equator, occurring about June 21 and December 22. sondage (French) A test excavation or test pit made at an archaeological site to determine the content and/or the distribution of prehistoric materials. Southeastern Ceremonial Complex A network of interaction, exchange, and shared information present over much of the southeastern (and parts of the midwestern) United States from around A.D. 1200 until the early 1500s; also previously referred to as the Southern Cult. spindle whorl A cam or balance wheel on a shaft or spindle for spinning yarn or thread from wool, cotton, or other material; usually made of clay. split inheritance An Andean practice by which the successor to the throne inherited only the office of the dead ruler; his junior kinsmen received the lands, palace, and personal wealth of the dead ruler. state A form of government with an internally specialized and hierarchically organized decision-making apparatus. A state generally has three or more administrative levels. status differentiation Inequality in human society in which certain individuals or groups have access to more resources, power, and roles than others. Differentiation occurs through ranking of descent groups or the creation of classes of people. steatite Soapstone, a variety of talc with a soapy or greasy feel; often used to make containers or carved ornaments. stela (Latin, pl. stelae) An erect stone monument that is often carved. stingray spine Bony tail spines of stingrays that were used in the past to draw blood in human autosacrificial rites. stirrup spout A distinctive curving spout on pottery vessels that is shaped like the stirrup of a saddle; characteristic of Moche pottery. stone boiling The process of heating stones in a fire and then adding them to containers to boil water or cook other foods.
pri31057_glo.ppG1-G10.qxd
10/2/09
8:28 AM
Page G-9
stratigraphic section The excavation of trenches and squares across manmade layers to expose a cross section of the deposits and reveal the sequence and methods of construction. stucco A type of plaster, often made out of lime, used for decoration. survey A systematic search of the landscape for artifacts and sites on the ground through aerial photography, field walking, soil analysis, and geophysical prospecting. sweat bath A hut or other space heated by steam that is created by pouring water over hot stones. Used by many peoples for ritual cleansing and therapeutic sweating. talud-tablero (Spanish) An architectural style characteristic of Teotihuacan during the Classic period, in which recessed rectangular panels (the tablero) are separated by sloping aprons (the talud ). tampu A roadside lodging and storage place (principally for food, fodder, firewood, and other commodities) along the Inca road system, placed roughly one day’s walk apart. technology The combination of knowledge and manufacturing techniques that enables people to convert raw materials into finished products. tell A mound composed of mud bricks and refuse, accumulated as a result of human activity. The mound of Jericho built up at a rate of roughly 26 cm (10 in) per 100 years, almost a foot per century. temper A nonplastic material (such as sand, shell, or fiber) that is added to clay to improve its workability and to reduce breakage during drying and firing. temporal marker A morphological type, such as a design motif on pottery or a particular type of stone tool, that has been shown to have a discrete and definable temporal range. teosinte (Aztec teo¯centli) A tall annual grass, native to Mexico and Central America, that is the closest relative of maize. terracotta A hard, brown-orange earthenware clay of fine quality, often used for architectural decorations, figurines, etc. tholoi Ancient Mesopotamian round structures that often were attached to a rectangular antechamber or annex, resulting in a keyhole shape. They
may have been used as storage facilities or as religious features for the interment of important individuals. tholos A large, beehive-shaped tomb, constructed using the corbel arch technique, characteristic of the Mycenaean civilization of Greece. tlachtli The Aztec word for their ritual ballgame. tool Any equipment, weapon, or object made by humans to change their environment. total station A computerized surveying and mapping instrument that uses a laser beam or radio waves to measure the distance and angle between the instrument and the target, and then calculates the exact position of the target. totem pole A pole or post that has been carved and painted with totems or figures, such as animals, that serve as the emblems of clans or families. Native Americans of the Pacific Northwest often erected these poles in front of their houses. transhumance A pattern of seasonal movement usually associated with pastoralists who take their herds to the mountains in summer and to the valleys in winter; more generally, a regular pattern of seasonal movement by human groups. trilithon A massive stone lintel occurring in prehistoric structures, such as Stonehenge and the tholos tombs in Greece. trophic level An organism’s place in the food chain. tuber A fleshy, usually oblong or rounded outgrowth (such as the potato) of a subterranean stem or root of a plant. tumpline A strap that is passed over the forehead or the chest to facilitate the transportation of a heavy load carried on the back. tzompantli The Aztec word for skull rack. The Aztec and other Mesoamerican peoples often placed the skulls of sacrificial victims on a wooden pole or frame; in some cases, large blocks of stone were sculpted to look like skull racks. UNESCO World Heritage Site A property around the world considered by the World Heritage Committee to have outstanding universal value. UNESCO (United Nations Educational, Scientific and Cultural
Organization) encourages the protection and preservation of cultural and natural heritage that is considered to be of outstanding value to humanity. unifacial A term describing a flaked stone tool in which only one face or side is retouched to make a sharp edge. vallum A Roman wall-and-ditch fortification. vault An arched structure of masonry that forms a ceiling or roof. The construction of vaults with corbelled, or stepped, ceilings was a common building technique of the Maya. wadi (Arabic) A dry streambed. waranqa A subdivision of the Inca empire that was used for administrative purposes, consisting of 1000 taxpayers. wattle and daub A building technique that uses a framework of poles, interspersed with smaller poles and twigs; the wooden frame is plastered with mud or a mud mixture. This building technique also was employed in the Southeast and other parts of the world. weaning The process of transferring the young from dependence on its mother ’s milk to other forms of nourishment. were-jaguar A representation of a supernatural figure that is half jaguar and half human, a common symbol in Preclassic Mesoamerica. wet-site excavation The technique of excavating waterlogged sites by pumping water through garden hoses to spray the dirt away and expose archaeological features and artifacts. wheel-thrown pottery Pottery that is made using the potter’s wheel. woodhenge A circular feature demarcated by large upright timbers; probably used by prehistoric groups as astronomical observatories. Zapotec A Mesoamerican cultural tradition generally associated with the Valley of Oaxaca and several smaller adjacent valleys in central Oaxaca (state of Oaxaca, Mexico). The Zapotec language is part of the Otomanguean language family, a language family that is distinct from the Maya or Utoaztecan, which includes Nahuatl, the language of the Aztecs. ziggurat A large pyramid in Mesopotamia consisting of many stepped levels. zoomorphic Having animal form or attributes.
_ Glossary
G-9
pri31057_glo.ppG1-G10.qxd
_ _
10/2/09
8:28 AM
Page G-10
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
Page R-1
<
References
Aaris-Sørensen, K., and E. Brinch Petersen. 1986. The Prejlerup aurochs— an archaeozoological discovery from Boreal, Denmark. Striae 24:111–117. Abbott, D. R., S. L. Stinson, and S. van Keuren. 2001. The economic implications of Hohokam buff ware exchange during the early Sedentary period. Kiva 67:7–29. Acker, R. 1998. New geographical tests of the hydraulic thesis at Angkor. South East Asia Research 6(1):5–47. Adams, D. 1980. The hitchhiker’s guide to the galaxy. New York: Harmony. Adams, J. L. 2002. Ground stone analysis: A technological approach. Salt Lake City: University of Utah Press. Adams, R. E. W. 1991. Prehistoric Mesoamerica, rev. ed. Norman: University of Oklahoma Press. Adams, R. E. W., W. E. Brown, and T. P. Culbert. 1981. Radar mapping, archeology and ancient Maya land use. Science 213:1457–1463. Adams, R. E. W., and M. J. MacLeod, eds. 2000. The Cambridge history of the native peoples of the Americas. Vol. 2, Mesoamerica, part 1. Cambridge: Cambridge University Press. Adams, R. McC. 1966. The evolution of urban society. Chicago: Aldine. Adams, R. McC. 1981. Heartland of cities. Chicago: University of Chicago Press. Adler, D. S., G. Bar-Oz, A. Belfer-Cohen, and O. Bar-Yosef. 2006. Ahead of the game: Middle and Upper Paleolithic hunting behaviors in the southern Caucasus. Current Anthropology 47:89–118. Aiello, L. C. 1993. The fossil evidence for modern human origins in Africa: A revised view. American Anthropologist 95:73–96. Aikens, C. M., and T. Higuchi. 1982. The prehistory of Japan. New York: Academic Press. Aikens, R. J. C. 1956. Stonehenge. Baltimore: Pelican. Aitken, M. J. 1985. Thermoluminescence dating. New York: Academic Press. Aitken, M. J. 1990. Science-based dating in archaeology. New York: Longman.
Akazawa, T. 1980. Fishing adaptation of prehistoric hunter-gatherers at the Nittano site, Japan. Journal of Archaeological Science 7:325–344. Alcock, S. E., T. N. D’Altroy, K. D. Morrison, and C. M. Sinopoli, eds. 2001. Empires: Perspectives from archaeology and history. Cambridge: Cambridge University Press. Alcock, S. E., and R. G. Osborne. 2005. Classical archaeology. Oxford: Blackwell. Algaze, G. 2001. Initial social complexity in southwestern Asia: The Mesopotamian advantage. Current Anthropology 42:199–233. Allchin, B., and R. Allchin. 1982. The rise of civilization in India and Pakistan. Cambridge: Cambridge University Press. Alperson-Afil, N., D. Richter, and N. Goren-Inbar. 2007. Phantom hearths and the use of fire at Gesher Benot Ya`Aqov, Israel. PaleoAnthropology 2007:1–15. Alva, W. 1990. New tomb of royal splendor: The Moche of ancient Peru. National Geographic 177(6):2–15. Alva, W. 2001. The royal tombs of Sipán: Art and power in Moche society. In Moche: Art and archaeology in ancient Peru, ed. J. Pillsbury. Studies in History of Art 63. Washington, DC: National Gallery of Art. Alva, W., and C. B. Donnan. 1993. Royal tombs of Sipán. Los Angeles: Fowler Museum of Cultural History, University of California. Alvarado Tezozomoc, F. 1975. Crónica Mexicáyotl, trans. E. O’Gorman. Universidad Nacional Autónoma de México, Mexico City: Originally written 1609. Ames, K. M. 1981. The evolution of social ranking on the Northwest Coast of North America. American Antiquity 46:789–805. Ammerman, A. J., and L. L. CavalliSforza. 1984. The Neolithic transition and the genetics of populations in Europe. Princeton, NJ: Princeton University Press. Anawalt, P. R. 1982. Understanding Aztec human sacrifice. Archaeology 35(3):38–45.
Anderson, A. 1987. Recent developments in Japanese prehistory: A review. American Antiquity 61:270–281. Anderson, D. G. and M. K. Faught. 1998. The distribution of fluted Paleoindian projectile points: Update 1998. Archaeology of Eastern North America 26:163–187. Anderson, D. G., and M. K. Faugh. 2000. Palaeoindian artefact distributions: Evidence and implications. Antiquity 74:507–513. Appenzeller, T. 1994. Clashing Maya superpowers emerge from a new analysis. Science 226:733–734. Arens, W. 1979. The man-eating myth: Anthropology and anthropophagy. New York: Oxford University Press. Arnold, B., and D. B. Gibson, eds. 1998. Celtic chiefdom, Celtic state. Cambridge: Cambridge University Press. Arnold, D. E., J. R. Branden, P. R. Williams, G. M. Feinman, and J. P. Brown. 2008. The first direct evidence for the production of Maya Blue: Rediscovery of a technology. Antiquity 82:151–164. Arsuaga, J. L., J. M. Carretero, A. Gracia, and I. Martínez. 1990. Taphonomical analysis of the human sample from the Sima de los Huesos Middle Pleistocene site (Atapuerca/Ibeas, Spain). Human Evolution 5:505–513. Arsuaga, J. L., I. Martínez, A. Gracia, J. M. Carretero, and E. Carbonell. 1993. Three new human skulls from the Sima de los Huesos Middle Paleolithic site in Sierra de Atapuerca, Spain. Nature 362:534–537. Arsuaga, J. L., I. Martínez, C. Lorenzo, A. Gracia, A. Muñoz, O. Alonso, and J. Gallego. 1999. The human cranial remains from Gran Dolina Lower Pleistocene site (Sierra de Atapuerca, Spain). Journal of Human Evolution 37:431–457. Atwood, Roger. 2004. Stealing history: Tomb raiders, smugglers, and the looting of the ancient world. New York: St. Martin’s Press. Aveni, A. F. 1986. The Nazca lines: Patterns in the desert. Archaeology 39(4):32–39.
References
R-1
_
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
Bahn, P. G., ed. 1995. The story of archaeology: The 100 great discoveries. New York: Barnes & Noble. Bahn, P. G., ed. 1996. The Cambridge illustrated history of archaeology. Cambridge: Cambridge University Press. Bahn, P. G. 1999. Bluff your way in archaeology. London: Oval Books. Bahn, P. G., ed. 2003. Written in bones: How human remains unlock the secrets of the dead. Toronto: Firefly Books. Bahn, P. G,. and J. Vertut. 1997. Journey through the Ice Age: Art and architecture. Berkeley: University of California Press. Bailey, G., and P. Spikins, eds. 2008. Mesolithic Europe. Cambridge: Cambridge University Press. Baines, J., and J. Málek. 1980. Atlas of ancient Egypt. New York: Facts on File. Balme, J., and A. Paterson. 2005. Archaeology in practice. A student guide to archaeological analyses. Oxford: Blackwell. Balter, V., J. Blichert-Toft, J. Braga, P. Telouk, F. Thackeray, and F. Albarède. 2008. U-Pb dating of fossil enamel from the Swartkrans Pleistocene hominid site, South Africa. Earth and Planetary Science Letters 267:236–246. Banning, E. B., and B. F. Byrd. 1987. Houses and changing residential units: Domestic architecture at PPNB ’Ain Ghazal, Jordan. Proceedings of the Prehistoric Society 53:8–65. Barber, R. L. N. 1988. The Cyclades in the Bronze Age. Iowa City: University of Iowa Press. Bareis, C. J., and J. W. Porter, eds. 1984. American Bottom archaeology. Urbana: University of Illinois Press. Barker, G. 1985. Prehistoric farming in Europe. Cambridge: Cambridge University Press. Bar-Yosef, O. 1986. The walls of Jericho: An alternative explanation. Current Anthropology 27:157–162. Bar-Yosef, O. 1998. The Natufian culture in the Levant: Threshold to the origins of agriculture. Evolutionary Anthropology 6:159–177. Bar-Yosef, O., and A. Belfer Cohen. 1992. Foraging to farming in the Mediterranean Levant. In Transitions to agriculture in prehistory, ed. A. B. Gebauer and T. D. Price. Madison, WI: Prehistory Press. Basu, S., J. Dickhaut, G. Hecht, K. Towry, and G. Waymire. 2009. Recordkeeping alters economic history by pro-
R-2
References
Page R-2
moting reciprocity. Proceedings of the National Academy of Sciences 106:1009–1014. Bauer, B. S., and R. A. Covey. 2002. Processes of state formation in the Inca heartland (Cuzco, Peru). American Anthropologist 104:846–864. Bawden, G. 1996. The Moche. Malden, MA: Blackwell. Bayard, D. 1971. Non Nok Tha: The 1968 excavation procedure, stratigraphy, and a summary of evidence. Studies in Prehistoric Anthropology, Vol. 4. Dunedin, NZ: University of Otago. Bayard, D. 1980. East Asia in the Bronze Age. In The Cambridge encyclopedia of archaeology, ed. A. Sherratt. New York: Crown. Beadle, G. 1980. The ancestry of corn. Scientific American 242:112–119. Becker, M. J. 1979. Priests, peasants, and ceremonial centers: The intellectual history of a model. In Maya archaeology and ethnohistory, ed. N. Hammond and G. R. Willey. Austin: University of Texas Press. Bellwood, P. 1978. Man’s conquest of the Pacific. Oxford: Oxford University Press. Bellwood, P. 1990. Foraging towards farming: A decisive transition or a millennial blur? Review of Archaeology 11:14–24. Bellwood, P. 2005. First farmers: The origins of agricultural societies. Malden, MA: Blackwell. Bender, B. 1978. Gatherer-hunter to farmer: A social perspective. World Archaeology 10:204–222. Bennett, W. C. 1934. Excavations at Tiahuanaco. Anthropological Papers of the American Museum of Natural History 34(3):359–494. Bennett, W. C. 1947. The archaeology of the central Andes. In Handbook of South American Indians. Vol. 2, The Andean civilizations, ed. J. Steward. Bureau of American Ethnology, Bulletin 143. Washington, DC: Smithsonian Institution. Benson, E. P., ed. 1968. Dumbarton Oaks conference on the Olmec. Washington, DC: Dumbarton Oaks. Benson, E. P., ed. 1971. Dumbarton Oaks conference on Chavín. Washington, DC: Dumbarton Oaks. Benson, E. P., ed. 1981. The Olmec and their neighbors: Essays in memory of Matthew W. Stirling. Washington, DC: Dumbarton Oaks.
Berdan, F. 1982. The Aztecs of central Mexico: An imperial society. New York: Holt, Rinehart & Winston. Berger, R., R. Chohfi, A. V. Zegarra, W. Yepez, and O. F. Carrasco. 1988. Radiocarbon dating Machu Picchu, Peru. Antiquity 62:707–710. Berlo, J. C., ed. 1992. Art, ideology, and the city of Teotihuacan. Washington, DC: Dumbarton Oaks. Bermudéz de Castro, J. M. A. 1998. Hominids at Atapuerca: The first human occupation in Europe. In The first Europeans: Recent discoveries and current debate, ed. E. Carbonell, J. Bermudéz de Castro, J. L. Arsuaga, and X. P. Rodriguez. Burgos, Spain: Aldecoa. Bernal, I. 1965. Archaeological synthesis of Oaxaca. In Handbook of Middle American Indians. Vol. 3, Archaeology of southern Mesoamerica, ed. G. R. Willey. Austin: University of Texas Press. Bernal, I. 1980. A history of Mexican archaeology: The vanished civilizations of Middle America. London: Thames & Hudson. Bicchieri, M. G. 1972. Hunters and gatherers today. New York: Holt, Rinehart & Winston. Bickerton, D. 1991. Language and species. Chicago: University of Chicago Press. Bidwell, P. T. 1985. The Roman fort of Vindolanda. HBMCE. Arch, Rep 1. London. Binford, L. R. 1968. Post-Pleistocene adaptations. In New perspectives in archeology, ed. S. R. Binford and L. R. Binford. Chicago: Aldine. Binford, L. R. 1983. In pursuit of the past. New York: Thames & Hudson. Binford, L. R. 2001. Constructing frames of reference: An analytical method for archaeological theory building using hunter-gatherer and environmental data sets. Berkeley: University of California Press. Binford, L. R., and S. R. Binford. 1966. A preliminary analysis of functional variability in the Mousterian of Levallois facies. In Recent studies in paleoanthropology, ed. J. D. Clark and F. C. Howell. American Anthropologist, special issue 68(2):238–295. Binford, L. R., and C. K. Ho. 1985. Taphonomy at a distance: Zhoukoudien, the cave home of Beijing man. Current Anthropology 26:413–442. Bingham, H. 1915. The story of Machu Picchu: The Peruvian expeditions of
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
the National Geographic Society and Yale University. National Geographic 27(2):172–216. Bingham, H. 1948. Lost city of the Incas. New York: Duell, Sloan & Pearce. Black, D. 1931. On an adolescent skull of Sinanthropus pekinensis in comparison with an adult skull of the same species and with other hominid skulls, recent and fossil. Palaeontologica Sinica, Series D, Vol. 7, Fasicule 2. Blanton, R. E. 1978. Monte Albán: Settlement patterns at the ancient Zapotec capital. New York: Academic Press. Blanton, R. E. 1983. The ecological perspective in highland Mesoamerican archaeology. In Archaeological hammers and theories, ed. J. A. Moore and A. S. Keene. New York: Academic Press. Blanton, R. E., G. M. Feinman, S. A. Kowalewski, and L. M. Nicholas. 1999. Ancient Oaxaca. Cambridge: Cambridge University Press. Blanton, R. E., S. A. Kowalewski, G. M. Feinman, and L. M. Finsten. 1993. Ancient Mesoamerica: A comparison of change in three regions. 2d ed. Cambridge: Cambridge University Press. Blumenschine, R. J. 1987. Characteristics of the early hominid scavenging niche. Current Anthropology 28:383–407. Bocherens, H. 2001. New isotopic evidence for dietary habits of Neanderthals from Belgium. Journal of Human Evolution 40:497–505. Bocquet, A., J. L. Brochier, A. EmeryBarbier, K. Lundstrom-Baudais, C. Orcel, and F. Vin. 1987. A submerged Neolithic village: Charavines “Les Baigneurs” in Lake Paladru, France. In European wetlands in prehistory, ed. J. M. Coles and A. J. Lawson. Oxford: Clarendon Press. Bogucki, P. 1988. Forest farmers and stockherders. Cambridge: Cambridge University Press. Bolger, D. Gender and human evolution. In Handbook of gender in archaeology, ed. S. M. Nelson. London: AltaMira. Bonnichsen, R., ed. 2004. Who were the first Americans? Corvallis, OR: Center for the Study of the First Americans. Bordaz, J. 1971. Tools of the Old and New Stone Age. New York: American Museum of Natural History.
Page R-3
Bordes, F. 1968. The Old Stone Age. New York: McGraw-Hill. Bordes, F. 1972. A tale of two caves. New York: Harper & Row. Bordes, F., and D. de Sonneville-Bordes. 1970. The significance of variability in Paleolithic assemblages. World Archaeology 2:61–73. Boserup, E. 1965. The conditions of agricultural growth: The economics of agrarian change under population pressure. Chicago: Aldine. Boule, M. 1911–1913. L’homme fossile de La Chapelle-aux-Saintes. Annales de Paléontologie, VI–VIII. Bourget, S. 2001. Rituals of sacrifice: Its practice at Huaca de la Luna and its representation in Moche iconography. In Moche art and archaeology in ancient Peru, ed. J. Pillsbury. New Haven, CT: National Gallery of Art. Bowen, D. Q. 1978. Quaternary geology. Oxford: Pergamon Press. Bradley, R. 1984. The social foundations of prehistoric Britain. Harlow, England: Longman. Bradley, R. 1998. The significance of monuments. London: Routledge. Braidwood, R. J. 1960. The agricultural revolution. Scientific American 203(3):130–148. Brain, C. K. 1981. The hunters or the hunted? An introduction to African cave taphonomy. Chicago: University of Chicago Press. Braudel, F. 1970. History and the social sciences: The long term. Social Science Information 9:145–174. Breeze, D., and B. Dobson. 2000. Hadrian’s Wall, 3d ed. London: Allen Lane. Brewer, D. J., and E. Teeter. 1999. Egypt and the Egyptians. Cambridge: Cambridge University Press. Brier, B. 2007. How to build a pyramid. Archaeology 60(3):22–27. Brodie, N., and K. W. Tubb, eds. 2002. Illicit antiquities: The theft of culture and the extinction of archaeology. London: Routledge. Brose, D., J. Brown, and D. Penney. 1985. Ancient art of the American Woodland Indians. New York: Harry N. Abrams. Brothwell, D. 1987. The bog man and the archaeology of people. Cambridge, MA: Harvard University Press. Brothwell, D., and A. M. Pollard, eds. 2001. Handbook of archaeological sciences. New York: Wiley.
Browman, D. L. 1981. New light on Andean Tiwanaku. American Scientist 69:408–419. Bruhns, K. O. 1994. Ancient South America. Cambridge: Cambridge University Press. Brumfiel, E. M., and G. M. Feinman, eds. 2008. The Aztec world. New York: Abrams. Brunhouse, R. L. 1973. In search of the Maya. New York: Ballantine Books. Bryan, A. L. 1978. Early man in America from a circum-Pacific perspective. Edmonton: Archaeological Researches International. Bryan, A. L. 1986. New evidence for the Pleistocene peopling of the Americas. Orono, ME: Center for the Study of Early Man. Bryant, V. 2007. Microscopic evidence for the domestication and spread of maize. Proceedings of the National Academy of Science 104:19,659–19,660. Buchanan, B., M. Collard, and K. Edinborough. 2008. Paleoindian demography and the extraterrestrial impact hypothesis. Proceedings of the National Academy of Science 105:11,651–11,654. Buikstra, J., and L. Beck. 2006. Bioarchaeology. The contextual analysis of human remains. San Diego: Elsevier. Burenhult, G. 1993. People of the Stone Age: Hunter-gatherers and early farmers. San Francisco: HarperCollins. Burenhult, G., ed. 1994. Old World civilizations: The rise of cities and states. New York: HarperCollins. Burenhult, G., ed. 1999. Arkeologi i Norden 1–2. Stockholm: Bokförlaget Natur och Kultur. Burger, R. L. 1984. The prehistoric occupation of Chavín de Huantar, Peru. Berkeley: University of California Press. Burger, R. L. 1985. Concluding remarks: Early Peruvian civilization and its relation to the Chavín horizon. In Early ceremonial architecture in the Andes, ed. C. B. Donnan. Washington, DC: Dumbarton Oaks. Burger, R. L. 1989. An overview of Peruvian archaeology (1976–1986). Annual Review of Anthropology 18:37–69. Burger, R. L. 1992. Chavín and the origins of Andean civilization. London: Thames & Hudson. Burger, R. L., and L. C. Salazar. 2004. Machu Picchu: Unveiling the mystery of the Incas. New Haven, CT: Yale University Press.
_ References
R-3
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
Burl, A. 1976. Stone circles of the British Isles. New Haven, CT: Yale University Press. Butzer, K. W. 1980. Civilizations: Organisms or systems? American Scientist 68:148–160. Butzer, K. W. 1982. Archaeology as human ecology. Cambridge: Cambridge University Press. Cabrera Castro, R., S. Sugiyama, and G. L. Cowgill. 1991. The Templo de Quetzalcoatl project at Teotihuacan. Ancient Mesoamerica 2:77–92. Calnek, E. E. 1976. The internal structure of Tenochtitlan. In The Valley of Mexico: Studies in pre-Hispanic ecology and society, ed. E. R. Wolf. Albuquerque: University of New Mexico Press. Campbell, B. G., and J. D. Loy. 2000. Humankind emerging, 8th ed. Boston: Longman. Cann, R. L., M. Stoneking, and A. C. Wilson. 1987. Mitochondrial DNA and human evolution. Nature 325:31–36. Carbonell, E., J. Castro, J. Pares, A. PerezGonzalez, G. Cuenca-Bescos, A. Olle, et al. 2008. The first hominin of Europe. Nature 452:465–469. Carman, J. 2002. Archaeology and heritage: An introduction. London: Continuum. Carneiro, R. L. 1970. A theory of the origin of the state. Science 169:733–738. Carneiro, R. L. 2003. Evolutionism in cultural anthropology: A critical history. Boulder, CO: Westview. Carr, C., and T. Case, eds. 2005. Gathering Hopewell: Society, ritual, and ritual interaction. New York: Kluwer. Carrasco, D., ed. 2001. The Oxford encyclopedia of Mesoamerican cultures: The civilizations of Mexico and Central America. Oxford: Oxford University Press. Carter, H., and A. C. Mace. 1923–1933. The tomb of Tut-ankh-Amen. London: Macmillan. Caso, A., and I. Bernal. 1965. Ceramics of Oaxaca. In Handbook of Middle American Indians. Vol. 3, Archaeology of southern Mesoamerica, ed. G. R. Willey. Austin: University of Texas Press. Cauvin, J. 2000. The birth of the gods and the origins of agriculture, trans. T. Watkins. Cambridge: Cambridge University Press. Cela-Conde, C. J., and F. J. Ayala. 2003. Genera of the human lineage. Proceedings of the National Academy of Sciences 100:7684–7689.
R-4
References
Page R-4
Chadwick, J. 1976. The Mycenaean world. Cambridge: Cambridge University Press. Chakrabarti, D. 1980. Early agriculture and the development of towns in India. In The Cambridge encyclopedia of archaeology, ed. A. Sherratt. New York: Crown. Chang, K. C. 1973. Food and food vessels in ancient China. Transactions of the New York Academy of Sciences 35:495–520. Chang, K. C. 1977a. Chinese archaeology since 1949. Journal of Asian Studies 36:623–646. Chang, K. C. 1977b. The continuing quest for China’s origins, I: Early farmers in China. Antiquity 30:116–123. Chang, K. C. 1977c. The continuing quest for China’s origins, II: The Shang civilization. Antiquity 30:187–193. Chang, K. C. 1981. In search of China’s beginnings: New light on an old civilization. American Scientist 69:148–160. Chang, K. C. 1986. The archaeology of ancient China. New Haven, CT: Yale University Press. Chang, K. C. 1989. Ancient China and its anthropological significance. In Archaeological thought in America, ed. C. C. Lamberg-Karlovsky. Cambridge: Cambridge University Press. Chang, K. C. 1994. Ritual and power. In Cradles of civilization: China, ed. R. E. Murowchick. Norman: University of Oklahoma Press. Changeux, J.-P., and J. Chavillon, eds. 1995. Origins of the human brain. Oxford: Oxford University Press. Chapman, R., I. Kinnes, and K. Randsborg, eds. 2009. The archaeology of death. Cambridge University Press. Chesterton, G. K. 1933. All I survey: A book of essays. London: Methuen. Childe, V. G. 1950. The urban revolution. Town Planning Review 21:3–17. Childe, V. G. 1951. Man makes himself. New York: New American Library. Childe, V. G. 1956. A short introduction to archaeology: Man and society. London: F. Muller. Chippendale, C. 1983. Stonehenge complete. Ithaca, NY: Cornell University Press. Churchill, S. E. 1998. Cold adaptation, heterochrony, and Neandertals. Evolutionary Anthropology 7:46–61. Cieza de León, P. 1959. The Incas, ed. V. von Hagen, trans. H. de Onis.
Norman: University of Oklahoma Press. Clark, J. D. 1970a. Kalambo Falls. Cambridge: Cambridge University Press. Clark, J. D. 1970b. The prehistory of Africa. London: Thames & Hudson. Clark, J. D., and S. A. Brandt, eds. 1984. From hunters to farmers. Berkeley: University of California Press. Clark, J. D., and J. W. K. Harris. 1985. Fire and its roles in early hominid lifeways. African Archaeological Review 3:3–28. Clark, J. E. 1986. From mountains to molehills: A critical review of Teotihuacan’s obsidian industry. In Research in Economic Anthropology, Supplement 2, ed. B. L. Isaac. Greenwich, CT: JAI Press. Clottes, J. 2008. Cave art. London: Phaidon. Clutton-Brock, J. 1999. A natural history of domesticated animals. Cambridge: Cambridge University Press. Coe, M. D. 1977. Mexico, 2d ed. New York: Praeger. Coe, M. D. 2005. The Maya, 7th ed. London: Thames & Hudson. Coe, M. D., and R. A. Diehl. 1980. In the land of the Olmec: The archaeology of San Lorenzo Tenochtitlan. Austin: University of Texas Press. Coe, M., D. Snow, and E. Benson. 1986. Atlas of ancient America. New York: Facts on File. Coe, W. R. 1965. Tikal: Ten years of study of a Maya ruin in the lowlands of Guatemala. Expedition 8(1):5–56. Coe, W. R. 1988. Tikal: A handbook of the ancient Maya ruins, 2d ed. Philadelphia: University Museum. Coe, W. R., and W. A. Haviland. 1982. Introduction to the archaeology of Tikal, Guatemala. University Museum Monograph 46. Philadelphia: University of Pennsylvania. Coggins, C. 1979. A new order and the role of the calendar: Some characteristics of the Middle Classic period at Tikal. In Maya archaeology and ethnohistory, ed. N. Hammond. Austin: University of Texas Press. Cohen, M. N. 1977. Population pressure and the origins of agriculture: An archaeological example from the coast of Peru. In The origins of agriculture, ed. C. Reed. The Hague: Mouton. Cole, S. 1975. Leakey’s luck: The life of Louis Seymour Bazett Leakey, 1903–1972. New York: Harcourt Brace Jovanovich.
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
Coles, J. M. 1982. The Bronze Age in northwestern Europe. Advances in World Archaeology 1:265–321. Coles, J. M., and E. S. Higgs. 1969. The archaeology of early man. London: Faber & Faber. Collis, J. 2001. Digging up the past: An introduction to archaeological excavation. Stroud, UK: Sutton. Conkey, M. W. 1980. The identification of prehistoric hunter-gatherer aggregation sites: The case of Altamira. Current Anthropology 21:609–630. Conkey, M. W. 1981. A century of Paleolithic cave art. Archaeology 34(4):20–28. Conrad, G. W. 1981. Cultural materialism, split inheritance, and the expansion of ancient Peruvian empires. American Antiquity 46:3–26. Conyers, L. B. 2004. Ground-penetrating radar for archaeology. Walnut Creek, CA: AltaMira. Coope, G. R. 1975. Climatic fluctuations in northwest Europe since the last interglacial, indicated by fossil assemblages of Coleoptera. In Ice ages: Ancient and modern, ed. A. E. Wright and F. Moseley. Liverpool: Seel House Press. Cordell, L. S. 1979. Prehistory: Eastern Anasazi. In Handbook of North American Indians. Vol. 9, Southwest, ed. A. Ortiz. Washington, DC: Smithsonian Institution Press. Cordell, L. S. 1997. Archaeology of the Southwest, 2d ed. San Diego: Academic Press. Cordell, L. S., and B. D. Smith. 1996. Indigenous farmers. In The Cambridge history of the native peoples of the Americas. Vol. 1, North America, part 1, ed. B. G. Trigger and W. E. Washburn. Cambridge: Cambridge University Press. Costantini, L. 1984. The beginning of agriculture in the Kachi Plain: The evidence of Mehrgarh. In South Asian archaeology 1981, ed. B. Allchin. New York: Cambridge University Press. Costin, C. L. 2004. Craft economies of ancient Andean states. In Archaeological perspectives on political economies, ed. G. M. Feinman and L. M. Nicholas. Salt Lake City: University of Utah Press. Cowan, C. W., and P. J. Watson. 1992. Origins of agriculture in world perspective. Washington, DC: Smithsonian Institution Press. Cowgill, G. L. 1975. Population pressure as a non-explanation. In Population
Page R-5
studies in archaeology and biological anthropology, ed. A. C. Swedlund. American Antiquity, Memoir 30:127–131. Cowgill, G. L. 1997. State and society at Teotihuacan, Mexico. Annual Review of Anthropology 26:129–161. Crawford, G. W., and C. Shen. 1998. The origins of rice agriculture: Recent progress in East Asia. Antiquity 72:858–866. Crawford, G. W., and H. Takamiya. 1990. The origins and implications of late prehistoric plant husbandry in northern Japan. Antiquity 64:889–911. Croes, D. R. 2003. Northwest Coast wetsite artifacts: A key to understanding resource procurement, storage, management, and exchange. In Emerging from the mist: Studies in Northwest Coast culture history, ed. R. G. Matson, G. Coupland, and Q. Mackie. Vancouver, BC: UBC Press. Crook, J. H. 1972. Sexual selection, dimorphism, and social organization in the primates. In Sexual selection and the descent of man, 1871–1971, ed. B. Campbell. Chicago: Aldine. Crown, P. L., and W. J. Judge, eds. 1991. Chaco and Hohokam: Prehistoric regional systems in the American Southwest. Santa Fe, NM: School of American Research Press. Crumley, C. L. 1995. Heterarchy and the anaysis of complex societies. In Hetararchy and the analysis of complex societies, ed. R. M. Ehrenreich, C. L. Crumley, and J. E. Levy. Archeological Papers No. 6. Arlington, VA: American Anthropological Association. Culbert, T. P. 1988. Political history and the Maya glyphs. Antiquity 62:135–152. Culbert, T. P., and D. S. Rice, eds. 1990. Precolumbian population history in the Maya Lowlands. Albuquerque: University of New Mexico Press. Cunliffe, B. 1994. The Oxford illustrated prehistory of Europe. Oxford: Oxford University Press. Cunliffe, B. 2001. Facing the ocean: The Atlantic and its people, 8000 B.C. to A.D. 1500. Oxford: Oxford University Press. Cunliffe, B. 2008. Europe between the oceans: 9000 B.C.–A.D. 1000. New Haven, CT: Yale University Press. Cunliffe, B., C. Gosden, and R. A. Joyce, eds. 2009. The Oxford handbook of archaeology. Oxford: Oxford University Press.
Curry, A. 2007. Digging into a desert mystery. Science 317:446–447. Dahlin, B. H. 1984. The colossus in Guatemala: The Preclassic Maya city of El Mirador. Archaeology 37(5):18–25. Dales, G. F. 1986. Some fresh approaches to old problems in Harappan archaeology. In Studies in the archaeology of India and Pakistan, ed. J. Jacobson. New Delhi: Oxford and IBH Publishing. D’Altroy, T. N. 1992. Provincial power in the Inka empire. Washington, DC: Smithsonian Institution Press. D’Altroy, T. N. 2001. Empires in a wider world. In Empires: Perspectives from archaeology and history, ed. S. E. Alcock, T. N. D’Altroy, K. D. Morrison, and C. M. Sinopoli. Cambridge: University of Cambridge Press. D’Altroy, T. N., and T. K. Earle. 1985. Staple finance, wealth finance, and storage in the Inka political economy. Current Anthropology 26:187–206. Dart, R. A. 1953. The predatory transition from ape to man. International Anthropological Linguistics Review 1:201–219. Darwin, C. 1981. The descent of man, and selection in relation to sex. With an introduction by J. Bonner and R. M. May. Princeton, NJ: Princeton University Press. Originally published 1871. Davidson, B. 1970. The lost cities of Africa, rev. ed. Boston: Little, Brown. Day, M. 1977. Guide to fossil man. London: Cassell. Deacon, H. 1989. Late Pleistocene paleoecology and archaeology in the southern Cape, South Africa. In The human revolution, ed. P. A. Mellars and C. B. Stringer. Princeton, NJ: Princeton University Press. Deacon, H. J., and J. Deacon. 1999. Human beginnings in South Africa: Uncovering the secrets of the Stone Age. Cape Town: David Philip. Dearborn, D. S. P., and K. J. Schreiber. 1986. Here comes the sun: The Cuzco–Machu Picchu connection. Archaeoastronomy 9:15–37. Dearborn, D. S. P., K. Schreiber, and R. E. White. 1987. Intimachay: A December solstice observatory at Machu Picchu, Peru. American Antiquity 52:346–352. de Borhegyi, S. F. 1980. The pre-Columbian ballgames: A pan-Mesoamerican tradition. Contributions in Anthropology and References
R-5
_
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
History: 1. Milwaukee: Milwaukee Public Museum. Decker-Walters, D., T. Walters, C. W. Cowan, and B. D. Smith. 1993. Isozymic characterization of wild populations of Cucurbita pepo. Journal of Ethnobiology 13:55–72. de Lumley, H. 1969. A Paleolithic camp at Nice. Scientific American 220(5):42–50. Demarest, A. 2004. Ancient Maya: The rise and fall of a rainforest civilization. Cambridge: Cambridge University Press. de Mortillet, G. 1872. Classification des ages de la pierre. Comptes rendues congress International d’Anthropologie et d’Archéologie prehistorique, VI session. Brussels. Denevan, W. M. 1992. The pristine myth: The landscape of the Americas in 1492. Annals of the Association of American Geographers 82:369–385. Denham, T., S. Haberle, and C. Lentfer. 2004. New evidence and revised interpretations of early agriculture in Highland New Guinea. Antiquity 78:839–857. Denham, T. P., S. G. Haberle, C. Lentfer, R. Fullagar, J. Field, M. Therin, N. Porch, and B. Winsborough. 2003. Origins of agriculture at Kuk Swamp in the highlands of New Guinea. Science 301:189–193. Dennell, R. C. 1983. European economic prehistory: A new approach. New York: Academic Press. d’Errico, F., C. Henshilwood, and P. Nilssen. 2001. An engraved bone fragment from ca. 70,000-year-old Middle Stone Age levels at Blombos Cave, South Africa: Implications for the origin of symbolism and language. Antiquity 75:309–318. Deuel, L. 1977. Memoirs of Heinrich Schliemann. New York: Harper & Row. Diamond, J. 1997. Guns, germs, and steel: The fates of human societies. New York: Norton. Diamond, J. 2002. Evolution, consequences and future of plant and animal domestication. Nature 418:700–707. Diáz del Castillo, B. 1956. The discovery and conquest of Mexico. New York: Farrar, Straus & Giroux. Diehl, R. A. 1976. Pre-Hispanic relationships between the Basin of Mexico and north and west Mexico. In The Valley of Mexico, ed. E. R. Wolf. Albu-
R-6
References
Page R-6
querque: University of New Mexico Press. Diehl, R. A. 1981. Tula. In Supplement to the handbook of Middle American Indians, ed. J. A. Sabloff. Austin: University of Texas Press. Diehl, R. A. 1983. Tula: The Toltec capital of ancient Mexico. London: Thames & Hudson. Diehl, R. A., and J. C. Berlo, eds. 1989. Mesoamerica after the decline of Teotihuacan, A.D. 700–900. Washington, DC: Dumbarton Oaks. Dikov, N. N. 1994. The Paleolithic of Kamchatka and Chukotka and the problem of the peopling of America. In Anthropology of the North Pacific Rim, ed. W. W. Fitzhugh and V. Chausronnet. Washington, DC: Smithsonian Institution Press. Dillehay, T. 1984. A late Ice-Age settlement in southern Chile. Scientific American 254(4):100–109. Dillehay, T. 1987. By the banks of the Chinchihuapi. Natural History 98(4):8–12. Dillehay, T. 1997. Monte Verde, a late Pleistocene settlement in Chile. Washington, DC: Smithsonian Institution Press. Dillehay, T. D. 2009. Probing deeper into first American studies. Proceedings of the National Academy of Science 106:971–978. Dillehay, T. D., J. Rossen, T. C. Andres, and D. E. Williams. 2008. Preceramic adoption of peanut, squash, and cotton in northern Peru. Science 316:1890–1893. Dinacauze, D. F. 2002. Environmental archaeology, principles and practice. Cambridge: Cambridge University Press. Dixon, J. E., J. R. Cann, and C. Renfrew. 1968. Obsidian and the origins of trade. Scientific American 211(3):44–53. Doebley, J. 1990. Molecular evidence and the evolution of maize. Economic Botany 44(3 Supplement):6–27. Domínguez-Rodrigo, M., and T. R. Pickering. 2003. Early hominid hunting and scavenging: A zooarchaeological review. Evolutionary Anthropology 12:275–282. Donnan, C. B. 1976. Moche art and iconography. Los Angeles: UCLA Latin American Center Publications. Donnan, C. B. 1990. Masterworks of art reveal a remarkable pre-Inca world. National Geographic 177(6):16–33.
Donnan, C. B., ed. 1985. Early ceremonial architecture in the Andes. Washington, DC: Dumbarton Oaks. Dorweiler, J., A. Stec, J. Kermicle, and J. Doebley. 1993. Teosinte glume architecture 1: A genetic locus controlling a key step in maize evolution. Science 262:233–235. Doyel, D. E., S. K. Fish, and P. R. Fish, eds. 2000. The Hohokam village revisited. Fort Collins, CO: Southwestern and Rocky Mountain Division of the American Association for the Advancement of Science. Drewett, P. 1999. Field archaeology: An introduction. London: Routledge. Drucker, P. 1955. Indians of the Northwest Coast. New York: McGraw-Hill. Drucker, P., R. Heizer, and R. Squier. 1959. Excavations at La Venta, Tabasco. Bureau of American Ethnology, Bulletin 170. Washington, DC: Smithsonian Institution. Dubois, E. 1894. Pithecanthropus erectus, eine Menschenahnliche Übergangsform aus Java. Cologne: Batavia. Duby, G. 1974. The early growth of the European economy: Warriors and peasants from the seventh to the twelfth century. Ithaca, NY: Cornell University Press. Dye, D. 1989. Death march of Hernando de Soto. Archaeology 42(3):27–31. Earle, T. 1997. How chiefs come to power: The political economy in prehistory. Stanford, CA: Stanford University Press. Elvin, M. 1973. The pattern of the Chinese past. Stanford, CA: Stanford University Press. Emerson, T. E., R. E. Hughes, M. R. Hynes, and S. U. Wisseman. 2003. The sourcing and interpretation of Cahokia-style figurines in the transMississippi South and Southeast. American Antiquity 68:287–313. Engel, F. A. 1976. An ancient world preserved. New York: Crown. English, N. B., J. L. Betancourt, J. S. Dean, and J. Quade. 2001. Strontium isotopes reveal distant sources of architectural timber in Chaco Canyon, New Mexico. Proceedings of the National Association of Science 98:11891–11896. Erickson, C. L. 1992. Prehistoric landscape management in the Andean highlands: Raised field agriculture and its environmental impact. Population and Environment 13:285–302. Erickson, D. L., B. D. Smith, A. C. Clarke, D. H. Sandweiss, and N. Tur-
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
oss. 2005. An Asian origin for a 10,000-year-old domesticated plant in the Americas. Proceedings of the National Academy of Sciences 102:18315–18320. Evans, D., C. Pottier, R. Fletcher, S. Hensley, I. Tapley, A. Milne, and M. Barbetti. 2007. A comprehensive archaeological map of the world’s largest preindustrial settlement complex at Angkor, Cambodia. Proceedings of the National Academy of Sciences 104(36):14,277–14,282. Evans, J., and T. O’Connor. 2001. Environmental archaeology, principles and method. Stroud, UK: Sutton. Evans, S. T. 2004. Ancient Mexico and Central America: Archaeology and culture history. London: Thames & Hudson. Evans, S. T., and D. L. Webster, eds. 2001. Archaeology of ancient Mexico and Central America: An encyclopedia. New York: Garland. Fagan, B. M. 1978. Quest for the past: Great discoveries in archaeology. Prospect Heights, IL: Waveland. Fagan, B. M. 1987. The great journey. London: Thames & Hudson. Fagan, B. M., ed. 1996. The Oxford companion to archaeology. Oxford: Oxford University Press. Fagan, B. M. 1998. 50 years of discovery. Archaeology 51(5):33–34. Fagan, B. M. 2000. Ancient North America, 3d ed. London: Thames & Hudson. Fagan, B. M. 2003. Archaeologists: Explorers of the human past. Oxford: Oxford University Press. Fagan, B. M. 2004. People of the earth: An introduction to world prehistory, 11th ed. Upper Saddle River, NJ: Prentice-Hall. Fagan, G. G., ed. 2006. Archaeological fantasies: How pseudoarchaeology misrepresents the past and misleads the public. London: Routledge. Fairservis, W. A. 1983. The script of the Indus Valley civilization. Scientific American 248(3):58–66. Falk, D. 1984. The petrified brain. Natural History 93(9):36–39. Farnsworth, P., J. E. Brady, M. J. deNiro, and R. S. MacNeish. 1985. A reevaluation of the isotopic and archaeological reconstructions of diet in the Tehuacán Valley. American Antiquity 50:102–116. Fash, W. L. 1991. Scribes, warriors, and kings: The city of Copán and the ancient Maya. London: Thames & Hudson.
Page R-7
Feder, K. L. 2001. Frauds, myths, and mysteries: Science and pseudoscience in archaeology, 4th ed. New York: McGraw-Hill. Feder, K. L., and M. A. Park. 2001. Human antiquity, 4th ed. New York: McGraw-Hill. Fedigan, L. M. 1986. The changing role of women in models of human evolution. Annual Review of Anthropology 15:25–66. Fedje, D. W., and H. Josenhans. 2000. Drowned forests and archaeology on the continental shelf of British Columbia, Canada. Geology 28:99–102. Feinman, G. M. 2001. Mesoamerican political complexity: The corporatenetwork dimension. In Leaders to rulers: The development of political centralization, ed. J. Haas. New York: Kluwer/Plenum. Feinman, G. M., S. A. Kowalewski, L. Finsten, R. E. Blanton, and L. M. Nicholas. 1985. Long-term demographic change: A perspective from the Valley of Oaxaca. Journal of Field Archaeology 12:333–362. Feinman, G. M., and L. Manzanilla, eds. 2000. Cultural evolution: Contemporary viewpoints. New York: Kluwer/Plenum. Feinman, G. M., and J. Marcus, eds. 1998. Archaic states. Santa Fe, NM: School for American Research Press. Feinman, G. M., and L. M. Nicholas. 2004. Unraveling the prehispanic highland Mesoamerican economy: Production, exchange, and consumption in the Classic period Valley of Oaxaca. In Archaeological perspectives on political eocnomies, ed. G. M. Feinman and L. M. Nicholas. Salt Lake City: University of Utah Press. Feinman, G. M., and T. D. Price, eds. 2001. Archaeology at the millennium: A sourcebook. New York: Kluwer/Plenum. Feldman, R. A. 1983. From maritime chiefdom to agricultural state in Formative coastal Peru. In Civilization in the ancient Americas: Essays in honor of Gordon R. Willey, ed. R. M. Leventhal and A. L. Kolata. Albuquerque: University of New Mexico Press. Fiedel, S. J. 1992. Prehistory of the Americas. Cambridge: Cambridge University Press. Findlayson, C. 2004. Neanderthals and modern humans: An ecological and evolutionary perspective. Cambridge: Cambridge University Press.
Finlayson, W. D. 1985. The 1975 and 1978 rescue excavations at the Draper site: Introduction and settlement patterns. National Museum of Man Mercury Series, Paper #130. Ottawa: Archaeological Survey of Canada. Finney, F. A., and J. B. Stoltman. 1991. The Fred Edwards site: A case of Stirling phase culture contact in southwestern Wisconsin. In New perspectives on Cahokia, ed. J. B. Stoltman. Madison, WI: Prehistory Press. Firestone, R. B., A. West, J. P. Kennett, L. Becker, T. E. Bunch, Z. S. Revay, P. H. Schultz, T. Belgya, D. J. Kennett, J. M. Erlandson, O. J. Dickenson, A. C. Goodyear, R. S. Harris, G. A. Howard, J. B. Kloosterman, P. Lechler, P. A. Mayewski, J. Montgomery, R. Poreda, T. Darrah, S. S. Que Hee, A. R. Smith, A. Sticr, W. Topping, J. H. Wittke, and W. S. Wolbach. 2007. Evidence for an extraterrestrial impact 12,900 years ago that contributed to the megafaunal extinctions and the Younger Dryas cooling. Proceedings of the National Academy of Science 104:16,016–16,021. Fish, P. R. 1998. Hohokam culture area. In Archaeology of prehistoric Native America: An encyclopedia, ed. G. Gibbon. New York: Garland. Fish, S. K., and P. R. Fish, eds. 2007. The Hohokam millennium. Santa Fe, NM: School for American Research Press. Fish, S. K., P. R. Fish, and J. H. Madsen, eds. 1992. The Marana community in the Hohokam world. Tucson: University of Arizona Press. Fish, S. K., and S. A. Kowalewski, eds. 1990. The archaeology of regions: A case for full-coverage survey. Washington, DC: Smithsonian Institution Press. Fisher, H. E. 1983. The sex contract: The evolution of human behavior. New York: Quill. Fitting, J. E. 1978. Regional cultural development, 300 B.C. to A.D. 1000. In Handbook of North American Indians. Vol. 15, Northeast, ed. W. C. Sturtevant and B. G. Trigger. Washington, DC: Smithsonian Institution Press. Flannery, K. V. 1968a. Archaeological systems theory and early Mesoamerica. In Anthropological archeology in the Americas, ed. B. J. Meggers. Washington, DC: Anthropological Society of Washington. Flannery, K. V. 1968b. The Olmec and the Valley of Oaxaca: A model for interregional interaction in Formative References
R-7
_
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
times. In Dumbarton Oaks conference on the Olmec, ed. E. Benson. Washington, DC: Dumbarton Oaks. Flannery, K. V. 1972a. The cultural evolution of civilizations. Annual Review of Ecology and Systematics 3:399–426. Flannery, K. V. 1972b. The origins of the village as a settlement type in Mesoamerica and the Near East: A comparative study. In Man, settlement, and urbanism, ed. P. J. Ucko, R. Tringham, and G. W. Dimbleby. London: Duckworth. Flannery, K. V. 1973. The origins of agriculture. Annual Review of Anthropology 2:271–310. Flannery, K. V., ed. 1976. The early Mesoamerican village. New York: Academic Press. Flannery, K. V., ed. 1986. Guilá Naquitz: Archaic foraging and early agriculture in Oaxaca, Mexico. New York: Academic Press. Flannery, K. V., and J. Marcus. 1976. Evolution of the public building in Formative Oaxaca. In Cultural change and continuity: Essays in honor of James Bennett Griffin, ed. C. Cleland. New York: Academic Press. Flannery, K. V., and J. Marcus. 1983. The growth of site hierarchies in the Valley of Oaxaca: Part 1. In The cloud people: Divergent evolution of the Zapotec and Mixtec civilizations, ed. K. V. Flannery and J. Marcus. New York: Academic Press. Flannery, K. V., and J. Marcus. 1994. Early Formative pottery of the Valley of Oaxaca, Mexico. Memoirs of the Museum of Anthropolog, No. 27. Ann Arbor: University of Michigan. Flannery, K. V., and J. Marcus, eds. 1983. The cloud people: Divergent evolution of the Zapotec and Mixtec civilizations. New York: Academic Press. Reprint, Clinton Corners, NY: Percheron Press, 2003. Flinders Petrie, W. M. 1904. Methods and aims of archaeology. London: Macmillan. Flint, R. F. 1971. Glacial and quaternary geology. New York: Wiley. Foley, R. 1987. Hominid species and stone-tool assemblages: How are they related? Antiquity 61:380–392. Foley, R. 1999. Evolutionary geography of Pliocene African hominids. In African biogeography, climate change, and human evolution, ed. T. G. Bromage and F. Schrenk. New York: Oxford University Press.
R-8
References
Page R-8
Folkens, P. A., and T. D. White. 2000. Human osteology. New York: Academic Press. Ford, J. A., and C. H. Webb. 1956. Poverty Point: A Late Archaic site in Louisiana. Anthropological Papers, vol. 46, part 1. New York: American Museum of Natural History. Ford, R. I., ed. 1984. The origins of plant husbandry in North America. Ann Arbor: University of Michigan Museum of Anthropology. Foster, J. 2003. Life and death in the Iron Age. Oxford: Ashmolean Museum Foster, M. S., and P. C. Weigand, eds. 1985. The archaeology of west and northwest Mesoamerica. Boulder, CO: Westview. Fowler, M. L. 1974. Cahokia: Ancient capital of the Midwest. Reading, MA: Addison-Wesley. Fowler, M. L. 1975. A pre-Columbian urban center on the Mississippi. Scientific American 233(2):92–101. Fowler, M. L. 1991. Mound 72 and Early Mississippian at Cahokia. In New perspectives on Cahokia, ed. J. B. Stoltman. Madison, WI: Prehistory Press. Fowler, M. L., and R. L. Hall. 1978. Late prehistory of the Illinois area. In Handbook of North American Indians. Vol. 15, Northeast, ed. W. C. Sturtevant and B. G. Trigger. Washington, DC: Smithsonian Institution Press. Frankfurt, H. 1956. The birth of civilization in the Near East. Garden City, NY: Doubleday. Frayer, D. W., M. H. Wolpoff, A. G. Thorne, F. H. Smith, and G. G. Pope. 1993. Theories of modern human origins: The paleontological test. American Anthropologist 95:14–50. French, E. 2002. Mycenae: Agamemnon’s capital. Stroud, UK: Tempus. Friedman, R. 2003. City of the hawk. Archaeology 56(6):50–56. Fullagar, R., J. Field, T. Denham, and C. Lentfer. 2006. Early and mid Holocene tool-use and processing of taro (Colocasia esculenta), yam (Dioscorea sp.) and other plants at Kuk Swamp in the highlands of Papua New Guinea. Journal of Archaeological Science 33:595–506. Funk, R. E. 1978. Post-Pleistocene adaptations. In Handbook of North American Indians. Vol. 15, Northeast, ed. W. C. Sturtevant and B. G. Trigger. Washington, DC: Smithsonian Institution Press.
Fyfe, C. 1994. The development of African states: 3000 B.C.–A.D. 1500. In Old World civilizations: The rise of cities and states, ed. G. Burenhult. San Francisco: HarperCollins. Galik, K., B. Senut, M. Pickford, D. Gommery, J. Treil, A. J. Kuperavage, and R. B. Eckhardt. 2004. External and internal morphology of the BAR 1002’00 Orrorin tugenensis femur. Science 305:1450–1452. Galinat, W. C. 1971. The origin of maize. Annual Review of Genetics 5:447–478. Gamble, C. 1986. The Paleolithic settlement of Europe. Cambridge: Cambridge University Press. Gamble, C. 1999. The Paleolithic societies of Europe. Cambridge: Cambridge University Press. Gamble, C. 2007. Origins and revolutions: Human identity in earliest prehistory. Cambridge: Cambridge University Press. Garlake, P. S. 1973. Great Zimbabwe. London: Thames & Hudson. Garlake, P. S. 1980. Early states in Africa. In The Cambridge encyclopedia of archaeology, ed. A. Sherratt. New York: Crown. Garrod, D. A. E., and D. M. A. Bate. 1937. The Stone Age of Mount Carmel. Oxford: Clarendon Press. Gates, C. 2003. Ancient cities. London: Routledge. Geertz, C. 1963. The transition to humanity. Anthropological Series 3:1–9. Washington, DC: Voice of America, United States Information Service. Gibbon, G., ed. 1998. Archaeology of prehistoric North America: An encyclopedia. New York: Garland. Gibson, J. L. 1987. The Poverty Point earthworks reconsidered. Mississippi Archaeology 22:15–31. Gibson, J. L. 1990. Earth sitting: Architectural masses at Poverty Point, northeastern Louisiana. In Recent research at the Poverty Point site, ed. K. M. Byrd. Lousiana Archaeology No. 13. Lafayette: Louisiana Archaeological Society. Gibson, J. L. 1996. Poverty Point: A terminal Archaic culture of the lower Mississippi Valley, 2d ed. Baton Rouge: Department of Culture, Recreation and Tourism, Louisiana Archaeological Survey and Antiquities Commission. Gibson, J. L. 2001. Ancient mounds of Poverty Point: Place of rings. Gainesville: University Press of Florida.
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
Gibson, J. L. 2006. Navels of the earth: Sedentism in the early moundbuilding cultures in the Lower Mississippi Valley. World Archaeology 38(2):311–329. Gilbert, W. H., and B. Asfaw. 2009. Homo erectus: Pleistocene evidence from the Middle Awash, Ethiopia. Berkeley: University of California Press. Gimbutas, M. 1977. Varna, a sensationally rich cemetery of the Karanova culture about 4500 B.C. Expedition 19, no. 4 (1977):39–47. Gingerich, P. D. 1985. Nonlinear molecular clocks and ape–human divergence times. In Hominid evolution: Past, present, and future, ed. P. V. Tobias. New York: A. R. Liss. Gleeson, P., and M. Fisken. 1977. Ozette archaeological project, interim final report, phase X. Pullman: Washington Archaeological Research Center, Washington State University. Gleeson, P., and G. Grosso. 1976. Ozette site. In The excavation of watersaturated archaeological sites (wet sites) on the Northwest Coast of North America, ed. D. R. Croes. Ottawa: Archaeological Survey of Canada. Glob, P. V. 1970a. The bog people. Ithaca, NY: Cornell University Press. Glob, P. V. 1970b. The mound people. Ithaca, NY: Cornell University Press. Glover, I. C. 1977. The Hoabinhian: Hunter-gatherers or early agriculturalists in Southeast Asia? In Hunters, gatherers, and first farmers beyond Europe, ed. J. V. S. Megaw. Leicester: Leicester University Press. Glover, I. C. 1980. Agricultural origins in East Asia. In The Cambridge encyclopedia of archaeology, ed. A. Sherratt. New York: Crown. Goebel, T., A. P. Derevianko, and V. T. Petrin. 1993. Dating the Middleto-Upper Paleolithic transition at Kara-Bom. Current Anthropology 34:452–458. Goebel, T., M. Waters, and D. O’Rourke. 2008. The Late Pleistocene dispersal of modern humans in the Americas. Science 319:1497–1502. Good, I. 2001. Archaeological textiles: A review of current research. Annual Review of Anthropology 30:209–226. Goodall, J. 1986. The chimpanzees of Gombe Reserve. Cambridge, MA: Harvard University Press. Goren-Inbar, N., N. Alperson, M. E. Kislev, O. Simchoni, Y. Melamed, A. Ben-Nun, and E. Werker. 2004.
Page R-9
Earliest signs of human-controlled fire uncovered in Israel. Science 5671:663–665. Gorman, C. H. 1970. Excavations at Spirit Cave, North Thailand: Some interim interpretations. Asian Perspectives 13:79–107. Gorman, C. H. 1971. The Hoabinhian and after: Subsistence patterns in Southeast Asia during the Late Pleistocene and Early Recent periods. World Archaeology 2:300–320. Gorman, C. H. 1977. A priori models and Thai prehistory: A reconsideration of the beginnings of agriculture in southeastern Asia. In The origins of agriculture, ed. C. A. Reed. The Hague: Mouton. Goudie, A. 1983. Environmental change. Oxford: Clarendon Press. Gould, S. J. 1984. A short way to corn. Natural History 93(3):12–20. Gowlett, J. A. J. 1984a. Ascent to civilization: The archaeology of early man. New York: Knopf. Gowlett, J. A. J. 1984b. Mental abilities of early man. In Community ecology and human adaptation in the Pleistocene, ed. R. A. Foley. London: Academic Press. Gowlett, J. A. J. 1987. The archaeology of accelerator radiocarbon dating. Journal of World Prehistory 1:127–170. Graham, I. 1967. Archaeological explorations in El Petén, Guatemala. Middle American Research Institute, Publication 33. New Orleans: Tulane University. Grayson, D. K. 1987. Death by natural causes. Natural History 96(5):8–12. Grayson, D. K. 1991. Late Pleistocene mammalian extinctions in North America: Taxonomy, chronology, and explanations. Journal of World Prehistory 5:193–232. Grayson, D. K. 1993. The desert’s past: A natural history of the Great Basin. Washington, DC: Smithsonian Institution Press. Grayson, D. K., and D. J. Meltzer. 2002. Clovis hunting and large mammal extinction: A critical review of the evidence. Journal of World Prehistory 16:313–359. Greber, N. B. 1998. Ohio Hopewell. In Archaeology of prehistoric Native America: An encyclopedia, ed. G. Gibbon. New York: Garland. Greber, N. B. 2003. Chronological relationships among Ohio Hopewell sites: Few dates and much complex-
ity. In Theory, method, and practice in modern archaeology, ed. R. J. Jeske and D. K. Charles. Westport, CT: Praeger. Green, M. W. 1981. The construction and implementation of the cuneiform writing system. Visible Language 15:345–372. Greene, K. 2002. Archaeology: An introduction, 4th ed. Philadelphia: University of Pennsylvania Press. Griffin, J. B. 1967. Eastern North American archaeology: A summary. Science 156:175–190. Griffin, J. B. 1980. Agricultural groups in North America. In The Cambridge encyclopedia of archaeology, ed. A. Sherratt. New York: Crown. Griffin, J. B. 1983. The Midlands. In Ancient North Americans, ed. J. Jennings. San Francisco: Freeman. Grove, D. C. 1981. The Formative period and the evolution of complex culture. In Supplement to the handbook of Middle American Indians, Vol. 1, ed. J. A. Sabloff. Austin: University of Texas Press. Grove, D. C. 1984. Chalcatzingo: Excavations on the Olmec frontier. London: Thames & Hudson. Grube, N., ed. 2001. Maya: Divine kings of the rain forest. Cologne: Könemann Verlagsgesellschaft mbH. Gumerman, G. J., ed. 1991. Exploring the Hohokam: Prehistoric desert peoples of the American Southwest. Dragoon, AZ: Amerind Foundation; Albuquerque: University of New Mexico Press. Gumerman, G. J., and E. W. Haury. 1979. Prehistory: Hohokam. In Handbook of North American Indians. Vol. 9, Southwest, ed. A. Ortiz. Washington, DC: Smithsonian Institution Press. Guthrie, R. D. 2005. The nature of Paleolithic art. Chicago: University of Chicago Press. Haas, J. S. 1982. The evolution of the prehistoric state. New York: Columbia University Press. Haas, J., ed. 2001. Leaders to rulers: The development of political centralization. New York: Kluwer/Plenum. Haas, J., W. Creamer, and A. Ruiz. 2004. Dating the Late Archaic occupation of the Norte Chico region in Peru. Nature 432:1020–1023. Haas, J., W. Creamer, and A. Ruiz. 2005. Power and the emergence of complex polities in the Peruvian preceramic. In Foundations of power in the prehispanic Andes, ed. K. J. Vaughn, References
R-9
_
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
D. Ogburn, and C. A. Conlee. Archeological Papers No. 14. Arlington, VA: American Anthropological Association. Haas, J., S. Pozorski, and T. Pozorski, eds. 1987. The origins and development of the Andean state. Cambridge: Cambridge University Press. Habu, J. 2004. Ancient Jomon of Japan. Cambridge: Cambridge University Press. Haddingham, E. 1979. Secrets of the Ice Age. London: Walker. Hall, M. 1996. Mapungubwe and Toutswemogala. In The Oxford companion to archaeology, ed. B. M. Fagan. New York: Oxford University Press. Hall, M., and S. Silliman. 2005. Historical archaeology. Malden, MA: Blackwell. Hall, R. L. 1977. An anthropocentric perspective for eastern United States prehistory. American Antiquity 42:499–518. Halloway, R. L. 1983. Cerebral brain endocast pattern of Australopithecus afarensis. Nature 303:420–422. Hammond, N. 1982. Ancient Maya civilization. New Brunswick, NJ: Rutgers University Press. Hammond, N. 1987. The discovery of Tikal. Archaeology 40(3):30–37. Hantman, J. L. 1990. Between Powhatan and Quirank: Reconstructing Monacan culture and history in the context of Jamestown. American Anthropologist 92:676–690. Hantman, J. L., and G. Dunham. 1993. The enlightened archaeologist. Archaeology 46(3):44–49. Harding, A. F. 2000. European societies in the Bronze Age. Cambridge: Cambridge University Press. Harlan, J. R. 1967. A wild wheat harvest in Turkey. Archaeology 20(3):197–201. Harlan, J. R. 1992. Crops and man, 2d ed. Madison, WI: American Society of Agronomy and Crop Science Society of America. Harlan, J. R. 1995. The living fields: Our agricultural heritage. Oxford: Oxford University Press. Harlan, J. R., J. M. J. de Wet, and A. B. L. Stemler, eds. 1976. Origins of African plant domestication. The Hague: Mouton. Harlan, J. R., and D. Zohary. 1966. Distribution of wild wheats and barley. Science 153:1074–1080. Harner, M. 1977. The enigma of Aztec sacrifice. Natural History 86(4):47–52.
R-10
References
Page R-10
Harris, D. R., and G. C. Hillman, eds. 1989. Foraging and farming: The evolution of plant exploitation. London: Unwin Hyman. Harrison, P. D. 1999. The lords of Tikal: Rulers of an ancient Maya city. London: Thames & Hudson. Harrison, R. J. 1980. The Beaker folk. London: Thames & Hudson. Hart, J. P., D. L. Asch, C. M. Scarry, and G. W. Crawford. 2002. The age of the common bean (Phaseolus vulgaris L.) in the northern Eastern Woodlands of North America. Antiquity 76:377–385. Harvati, K., and T. Harrison, eds. 2007. Neanderthals revisited: New approaches and perspectives. New York: Springer. Hassan, F. A. 1981. Demographic archaeology. New York: Academic Press. Hassan, F. A. 1997. Global population and human evolution. Human Evolution 12:3. Hassan, F. A. 2007. The lie of history: Nation-states and the contradictions of complex societies. In Sustainability or collapse: An integrated history and future of the people on Earth, ed. R. Costanza, L. J. Graumlich, and W. Steffen. Cambridge, MA: MIT Press. Hastings, C. M., and M. E. Moseley. 1975. The adobes of Huaca del Sol and Huaca de la Luna. American Antiquity 40:196–203. Hastorf, C. A., ed. 1999. Early settlement at Chiripa, Bolivia: Research of the Taraco archaeological project. Contributions of the Archaeological Research Facility No. 57. Berkeley: University of California. Hastorf, C. A., and V. S. Popper. 1989. Current paleoethnobotany. Chicago: University of Chicago Press. Haury, E. W. 1976. The Hohokam: Desert farmers and craftsmen. Tucson: University of Arizona Press. Hayden, B. 1990. Nimrods, piscators, pluckers and planters: The emergence of food production. Journal of Anthropological Archaeology 9:31–69. Haynes, G. 2009. American megafaunal extinctions at the end of the pleistocene. New York: Springer. Heckenberger, M. J., A. Kuikuro, U. T. Kuikuro, J. C. Russell, M. Schmidt, C. Fausto, and B. Franchetto. 2003. Amazonia 1492: Pristine forest or cultural parkland? Science 301:1710–1714.
Heckenberger, M. J., J. B. Petersen, and E. Goés Neves. 1999. Village size and permanence in Amazonia: Two archaeological examples from Brazil. Latin American Antiquity 10:353–376. Hedges, R. E. M. 1981. Radiocarbon dating with an accelerator. Archaeometry 23:3–18. Helbaek, H. 1960. The paleoethnobotany of the Near East and Europe. In Prehistoric investigations in Iraqi Kurdistan, ed. R. J. Braidwood and B. Howe. Studies in Oriental Civilization 31. Chicago: Oriental Institute. Henderson, J. S. 1981. The world of the ancient Maya. Ithaca, NY: Cornell University Press. Henke, W., and I. Tattersall. 2006. Handbook of paleoanthropology. London: Springer. Henry, D. 1989. From foraging to agriculture: The Levant at the end of the Ice Age. Philadelphia: University of Pennsylvania Press. Henshilwood, C. S., F. d’Errico, R. Yates, Z. Jacobs, C. Tribolo, G. A. T. Duller, N. Mercier, J. C. Sealy, H. Valladas, I. Watts, and A. Wintle. 2002. Emergence of modern human behaviour: Middle Stone Age engravings from South Africa. Science 295:1278–1280. Hesse, B. 1982. Slaughter patterns and domestication: The beginnings of pastoralism in western Iran. Man 17:403–417. Higham, C. F. W. 1977. Economic change in prehistoric Thailand. In The origins of agriculture, ed. C. A. Reed. The Hague: Mouton. Higham, C. F. W. 1984. Prehistoric rice cultivation in Southeast Asia. Scientific American 250(4):138–146. Higham, C. F. W. 2001. The civilization of Angkor. Berkeley: University of California Press. Higham, C. F. W. 2002. Early cultures of mainland Southeast Asia. Chicago: Art Media Resources. Higham, C. F. W., R. Bannanurag, G. Mason, and N. Tayles. 1992. Human biology, environment, and ritual at Khok Phanom Di. World Archaeology 24:35–54. Higham, C. F. W., and A. Kijngam. 1982. Prehistoric man and his environment: Evidence from the Ban Chiang faunal remains. Expedition 24(4):17–24. Higham, C. F. W., and T. L.-D. Lu. 1998. The origins and dispersal of rice cultivation. Antiquity 72:867–877.
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
Higham, C. F. W., and R. Thosarat. 1994. Khok Phanom Di: Prehistoric adaptation to the world’s richest habitat. Fort Worth: Harcourt Brace. Hill, B., and R. Hill. 1974. Indian petroglyphs of the Pacific Northwest. Saanichton, Canada: Hancock House. Hill, J. N. 1970. Broken K Pueblo: Prehistoric social organization in the American Southwest. Tucson: University of Arizona Press. Hillman, G. C., and M. S. Davies. 1990. Measured domestication rates in wild wheats and barley under primitive cultivation, and their archaeological implications. Journal of World Prehistory 4:157–222. Hodder, I. 2000. Towards reflexive method in archaeology: The example at Çatalhöyük. Cambridge: McDonald Institute for Archaeological Research. Hodder, I., ed. 2001. Archeological theory today. Cambridge: Polity Press. Hodder, I. 2006. The leopard’s tale: Revealing the mysteries of Çatalhöyük. New York: Thames & Hudson. Hoffecker, J. F., W. R. Powers, and T. Goebel. 1993. The colonization of Beringia and the peopling of the New World. Science 259:46–53. Hoffman, M. A. 1976. The city of the hawk. Expedition 18(3):32–41. Hoffman, M. A. 1983. Where nations began. Science 83:42–51. Hole, F., K. V. Flannery, and J. A. Neely. 1969. Prehistory and human ecology of the Deh Luran Plain. Ann Arbor: University of Michigan Press. Holliday, V. T. 2004. Soils in archaeological research. Oxford: Oxford University Press. Holloway, R. L. 1975. The role of human social behavior in the evolution of the brain. New York: American Museum of Natural History. Hood, S. 1973. The Minoans. London: Thames & Hudson. Hsu, C. 1965. Ancient China in transition. Stanford, CA: Stanford University Press. Huckell, B. B. 1996. The Archaic prehistory of the North American Southwest. Journal of World Prehistory 10:305–373. Hyslop, J. 1984. The Inka road system. New York: Academic Press. Ikawa-Smith, F. 1980. Current issues in Japanese archaeology. American Scientist 68:134–145.
Page R-11
Ikram, S. 2009. Ancient Egypt: An introduction. Cambridge: Cambridge University Press. Iltis, H. H. 1983. From teosinte to maize: The catastrophic sexual transmutation. Science 222:886–894. Isaac, G. 1977. Olorgesailie: Archaeological studies of a Middle Pleistocene lake basin in Kenya. Chicago: University of Chicago Press. Isaac, G. 1984. The archaeology of human origins: Studies of the Lower Pleistocene in East Africa, 1971–1981. Advances in World Archaeology 3:1–87. Isaac, G., and R. Leakey. 1979. Human ancestors. Readings from Scientific American. San Francisco: Freeman. Isbell, W. H. 1978. The prehistoric ground drawings of Peru. Scientific American 238(4):140–153. Isbell, W. H., and H. Silverman, eds. 2002a. Andean archaeology I: Variations in sociopolitical organization. New York: Kluwer/Plenum. Isbell, W. H., and H. Silverman, eds. 2002b. Andean archaeology II: Art, landscape, and society. New York: Kluwer/ Plenum. Jackson, H. E. 1989. Poverty Point adaptive systems in the lower Mississippi Valley: Subsistence remains from the J. W. Copes site. North American Archaeologist 10:173–203. Jackson, H. E. 1998. Poverty Point objects. In Archaeology of prehistoric Native America: An encyclopedia, ed. G. Gibbon. New York: Garland. Jacobsen, T. 1976. Seventeen thousand years of Greek prehistory. Scientific American 234(6):76–87. Jacobson, J. 1979. Recent developments in South Asian prehistory and protohistory. Annual Review of Anthropology 8:467–502. Jacobson, J. 1986. The Harappan civilization: An early state. In Studies in the archaeology of India and Pakistan, ed. J. Jacobson. New Delhi: Oxford and IBH Publishing. Jameson, J. H. 1997. Presenting archaeology to the public: Digging for truths. Walnut Creek, CA: AltaMira. Janusek, J. W. 2004. Tiwanaku and its precursors: Recent research and emerging perspectives. Journal of Archaeological Research 12:121–183. Jarrige, J.-F., and R. H. Meadow. 1980. The antecedents of civilization in the Indus Valley. Scientific American 243(2):122–133.
Jawad, A. J. 1974. The Eridu material and its implications. Sumer 30:11–46. Jefferson, T. 1797. Notes of the state of Virginia. London: J. Stockdale. Jeffries, R. W. 1987. The archaeology of Carrier Mills. Carbondale: Southern Illinois University Press. Jefferies, R. W. 2009. Holocene huntergatherers of the Lower Ohio River Valley. Tuscaloosa: University of Alabama Press. Jeffries, R. W., and M. Lynch. 1985. Dimensions of Middle Archaic cultural adaptation at the Black Earth site, Saline County, Illinois. In Archaic hunters and gatherers in the American Midwest, ed. J. L. Phillips and J. A. Brown. New York: Academic Press. Jelinek, A. J. 1982. The Tabun Cave and Paleolithic man in the Levant. Science 216:1369–1375. Jelinek, A. J. 1988. Technology, typology, and culture in the Middle Paleolithic. In Upper Pleistocene prehistory, ed. H. Dibble and A. Montet-White. Philadelphia: University of Pennsylvania Press. Jennings, J. D., ed. 1983. Ancient South Americans. San Francisco: Freeman. Jiang, L., and L. Liu. 2006. New evidence for the origins of sedentism and rice domestication in the Lower Yangzi River, China. Antiquity 80:355–361. Jiménez Moreno, W. 1941. Tula y los toltecas según las fuentes históricas. Revista Mexicana de Estudios Antropológicos 5:79–83. Joffroy, R. 1962. Le Trésor de Vix. Histoire et portée d’une grande découverte. Paris: Fayard. Johansen, K. L., S. T. Laursen, and M. K. Holst. 2004. Spatial patterns of social organization in the Early Bronze Age of South Scandinavia. Journal of Anthropological Archaeology 23:33–55. Johanson, D. C. 1976. Ethiopia yields first “family” of early man. National Geographic 150(6):790–811. Johanson, D. C., and M. A. Eddy. 1981. Lucy: The beginnings of humankind. New York: Simon & Schuster. Johnson, M. 1999. Archaeological theory. Oxford: Blackwell. Jolly, C. 1970. The seed eaters: A new model of hominid differentiation based on a baboon analogy. Man 5:5–26. Jones, A., ed. 2008. Prehistoric Europe: Theory and practice. Oxford: Blackwell. References
R-11
_
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
Jones, C. 1977. Inauguration dates of three Late Classic rulers of Tikal, Guatemala. American Antiquity 42:28–60. Jones, C., and L. Satterthwaite. 1982. The monuments and inscriptions of Tikal: The carved monuments. University Museum Monograph 44. Philadelphia: University of Pennsylvania. Jones, M. 2003. The molecule hunt: Archaeology and the search for ancient DNA. New York: Arcade. Jones, M. K., R. G. Allaby, T. A. Brown, F. Hole, M. Heun, B. Borghi, and F. Salamini. 1998. Wheat domestication. Science 279:202–204. Jurmain, R., H. Nelson, and W. A. Turnbaugh. 1987. Understanding physical anthropology and archaeology, 3d ed. St. Paul, MN: West. Kantner, J. 2004. Ancient Puebloan Southwest. Cambridge: Cambridge University Press. Kay, R. F., and F. E. Grine. 1988. Tooth morphology, wear and diet in Australopithecus and Paranthropus from southern Africa. In Evolutionary history of the “robust” Australopithecines, ed. F. E. Grine. New York: Aldine de Gruyter. Keatinge, R. W., ed. 1988. Peruvian prehistory. Cambridge: Cambridge University Press. Keeley, L. H. 1981. Experimental determination of stone tool uses: A microwear analysis. Chicago: University of Chicago Press. Keeley, L. H., and N. Toth. 1981. Microwear polishes on early stone tools from Koobi Fora, Kenya. Nature 293(8):464–465. Keightley, D. N., ed. 1983. The origins of Chinese civilization. Berkeley: University of California Press. Kelly, R. L. 1995. The foraging spectrum: Diversity in hunter-gatherer lifeways. Washington, DC: Smithsonian Institution Press. Kelly, R. L., and M. Prasciunas. 2004. Did the ancestors of Native Americans cause animal extinctions in Late Pleistocene North America? In Reconsidering the ecological Indian, ed. M. E. Harkin and D. R. Lewis. Lincoln: University of Nebraska Press. Kennedy, G. E. 2005. From the ape’s dilemma to the weanling’s dilemma: Early weaning and its evolutionary context. Journal of Human Evolution 48:123–145.
R-12
References
Page R-12
Kennett, D. J., J. P. Kennett, A. West, C. Mercer, S. S. Que Hee, L. Bement, T. E. Bunch, M. Sellers, and W. S. Wolbach. 2009. Nanodiamonds in the Younger Dryas Boundary Sediment Layer. Science 323:94. Kenoyer, J. M. 1984. Shell working industries of the Indus civilization: A summary. Paleorient 10(1):49–63. Kenoyer, J. M. 1985. Shell working at Mohenjo-daro, Pakistan. In South Asian archaeology 1983, ed. J. Schotsmans and M. Taddei. Naples: Instituto Universitario Orientale. Kenoyer, J. M. 1991. The Indus Valley tradition of Pakistan and western India. Journal of World Prehistory 5:331–385. Kenoyer, J. M. 1998a. Ancient cities of the Indus Valley civilization. Oxford: Oxford University Press. Kenoyer, J. M. 1998b. Birth of a civilization. Archaeology 51(1):54–61. Kense, F. J., and J. A. Okoro. 1993. Changing perspectives on traditional iron production in West Africa. In The archaeology of Africa: Food, metals and towns, ed. T. Shaw, P. Sinclair, B. Andah, and A. Okpoko. London: Routledge. Kenyon, K. 1954. Ancient Jericho. Scientific American 190(4):76–82. Kenyon, K. 1960. Excavations at Jericho. I. Jerusalem: British School of Archaeology. Kidder, T. R. 2008. Poverty Point and the archaeology of singularity. The SAA Archaeological Record 8(5):9–12. King, A. 2003. Over a century of explorations at Etowah. Journal of Archaeological Research 11(4):279–306. Kirchhoff, P. 1952. Mesoamerica: Its geographic limits, ethnic composition, and cultural characteristics. In Heritage of conquest, ed. S. Tax. New York: Free Press. Kirk, R., and R. D. Daugherty. 1978. Exploring Washington archaeology. Seattle: University of Washington Press. Kislev, M. E., A. Hartmann, and O. Bar-Yosef. 2006. Early domesticated fig in the Jordan Valley. Science 312:1372–1374. Kittler, R., M. Kayser, and M. Stoneking. 2003. Molecular evolution of Pediculus humanus and the origin of clothing. Current Biology 13:1414–1417. Klein, R. G. 1995. Anatomy, behavior, and modern human origins. Journal of World Prehistory 9:167–198.
Klein, R. G. 1999. The human career, 2d ed. Chicago: University of Chicago Press. Klein, R. G., and K. Cruz-Uribe. 1984. The analysis of animal bones from archaeological sites. Chicago: University of Chicago Press. Klein, R. G., and K. Cruz-Uribe. 1987. Large mammal and tortoise bones from Eland’s Bay Cave Province, South Africa. In Papers in the prehistory of the Western Cape, South Africa, ed. J. Parkington and M. Hall. Oxford: British Archaeological Reports. Klein, R. G., and B. Edgar. 2002. The dawn of human culture. New York: Wiley. Klima, B. 1962. The first ground plan of an Upper Paleolithic loess settlement in middle Europe and its meaning. In Courses toward urban life, ed. R. J. Braidwood and G. R. Willey. Chicago: Aldine. Klima, B. 1963. Dolni Vestonice. Prague: Nakladatelstvi Ceskoslovenske Akademie Ved. Knight, V. J., Jr. 1990. Social organization and the evolution of hierarchy in southeastern chiefdoms. Journal of Anthropological Research 46:1–23. Knight, V. J., Jr., and V. P. Steponaitis, eds. 1998. Archaeology of the Moundville chiefdom. Washington, DC: Smithsonian Institution Press. Kobayashi, T. 2004. Jomon reflections: Forager life and culture in the prehistoric Japanese archipelago. Oxford: Oxbow Books. Kojan, D. 2008. Paths of power and politics: Historical narratives at the Bolivian site of Tiwanaku. In Evaluating multiple narratives: Beyond nationalist, colonialist, imperialist archaeologies, ed. J. Habu, C. Fawcett, and J. M. Matsunaga. New York: Springer. Kolata, A. L. 1983. The South Andes. In Ancient South Americans, ed. J. D. Jennings. San Francisco: Freeman. Kolata, A. L. 1986. The agricultural foundations of the Tiwanaku state. American Antiquity 51:748–762. Kolata, A. L. 1987. Tiwanaku and its hinterland. Archaeology 40(1):36–41. Kolata, A. L. 1993. The Tiwanaku: Portrait of an Andean Civilization. Cambridge, MA: Blackwell. Kolata, A. L., ed. 2003. Tiwanaku and its hinterland: Archaeology and paleoecology of an Andean civilization. Washington, DC: Smithsonian Institution Press.
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
Kowalewski, S. A., G. M. Feinman, L. Finsten, R. E. Blanton, and L. M. Nicholas. 1989. Monte Albán’s hinterland, part II: Prehispanic settlement patterns in Tlacolula, Etla, and Ocotlán, the Valley of Oaxaca, Mexico. Memoirs of the Museum of Anthropology, No. 23. Ann Arbor: University of Michigan. Kowalski, J. K., and C. Kristan-Graham, eds. 2007. Twin Tollans: Chichén Itzá, Tula, and the Epiclassic to Early Postclassic Mesoamerican world. Washington, DC: Dumbarton Oaks. Kramer, S. N. 1988. The temple in Sumerian literature. In Temple in society, ed. M. V. Fox. Winona Lake, IN: Eisenbrauns. Krings, M., A. Stone, R. W. Schmitz, H. Krainitzki, M. Stoneking, and S. Pääbo. 1997. Neanderthal DNA sequences and the origin of modern humans. Cell 90:19–30. Kristiansen, K. 2000. Europe before history. The European world system in the 2nd millennium B.C. Cambridge: Cambridge University Press. Kuman, K., and R. J. Clarke. 2000. Stratigraphy, artefact industries and hominid associations for Sterkfontein, Member 5. Journal of Human Evolution 38:827–847. Kurtén, B. 1968. Pleistocene mammals of Europe. Chicago: Aldine. Kurtén, B., and E. Anderson. 1980. Pleistocene mammals of North America. New York: Columbia University Press. Kusimba, C. M. 2008. Early African cities: Their role in the shaping of urban and rural interaction spheres. In The ancient city: New perspectives on urbanism in the Old and New Worlds, ed. J. Marcus and J. A. Sabloff. Santa Fe, NM: School for Advanced Research Press. Laitman, J. T. 1984. The anatomy of human speech. Natural History 93(9):20–27. Lamberg-Karlovsky, C. C., and J. A. Sabloff. 1995. Ancient civilizations: The Near East and Mesoamerica, 2d ed. Prospect Heights, IL: Waveland Press. Lambert, J. 1997. Traces of the past: Unraveling the secrets of archaeology through chemistry. New York: Addison Wesley Longman. Lanning, E. P. 1967. Peru before the Incas. Englewood Cliffs, NJ: Prentice-Hall. Larsen, C. S. 1999. Bioarchaeology: Interpreting behavior from the human skeleton.
Page R-13
New York: Cambridge University Press. Larsen, C. S. 2000. Skeletons in our closet: Revealing our past through bioarchaeology. Princeton, NJ: Princeton University Press. Larsen, C. S., R. M. Matter, and D. L. Cabo. 1998. Human origins: The fossil record. Prospect Heights, IL: Waveland Press. Larsson, L. 1988. The Skateholm Project. I, Man and environment. Lund, Sweden: Almqvist & Wiksell International. Lawton, G. 2004. Urban legends. New Scientist (18 September):32–35. Leacock, E. B. 1971. North American Indians in historical perspective. New York: Random House. Leakey, M. D. 1971. Olduvai Gorge. Cambridge: Cambridge University Press. Leakey, M. D. 1978. Pliocene footprints at Laetoli, Tanzania. Antiquity 52:133. Leakey, M. D., and J. M. Harris, eds. 1987. Laetoli: A Pliocene site in northern Tanzania. Oxford: Clarendon Press. Leakey, M. D., and R. E. Leakey, eds. 1978. Koobi Fora research project. Oxford: Clarendon Press. Leakey, M. G., F. Spoor, F. H. Brown, P. N. Gathogo, C. Klarle, L. N. Leakey, and I. McDougall. 2001. New hominin genus from eastern Africa shows diverse middle Pliocene lineages. Nature 410:433–440. Leakey, R. 1981. The making of mankind. London: M. Joseph. Leakey, R., and R. Lewin. 1977. Origins reconsidered. New York: Dutton. Lechevallier, M., and G. Quivron. 1985. Results of the recent excavations at the Neolithic site of Mehrgarh, Pakistan. In South Asian archaeology 1983, ed. J. Schotsmans and M. Taddei. Naples: Instituto Universitario Orientale. Lee, R. B., and R. Daly, eds. 1999. Cambridge encyclopedia of hunters and gatherers. Cambridge: Cambridge University Press. Lee, R. B., and I. DeVore. 1968. Man the hunter. Chicago: Aldine. Legge, A. J., and P. A. Rowley-Conwy. 1988. Star Carr revisited. London: University of London. LeGros Clark, W. E., and B. G. Campbell. 1978. The fossil evidence for human evolution. Chicago: University of Chicago Press. Lekson, S. H. 1999. The Chaco meridian: Centers of political power in the an-
cient Southwest. Walnut Creek, CA: AltaMira. Lekson, S. H., T. C. Windes, J. R. Stein, and W. J. Judge. 1988. The Chaco Canyon community. Scientific American 259(1):72–81. León-Portilla, M. 1987. Ethnohistorical record for the Huey Teocalli. In The Aztec Templo Mayor, ed. E. H. Boone. Washington, DC: Dumbarton Oaks. Lepper, B. 2005. Ohio archaeology: An illustrated chronicle of Ohio’s ancient American Indian cultures. Wilmington, OH: Orange Frazer Press. Leroi-Gourhan, A. 1957. Prehistoric man. New York: Philosophical Library. Leroi-Gourhan, A. 1968. The archaeology of Lascaux Cave. Scientific American 219(4):104–111. Leroi-Gourhan, A. 1984. The dawn of European art: An introduction to Paleolithic cave paintings. Cambridge: Cambridge University Press. Leroi-Gourhan, A., and M. Brezillon. 1972. Fouilles de Pincevent: Essai d’analyse ethnographique d’un habitat Magdalenien (sec. 36). VII supplément à Gallia Prehistoria. Paris: Éditions du Centre National de la Recherche Scientifique. Levtzion, N. 1976. The early states of the western Sudan to 1500. In History of West Africa, vol. 1, 2d ed., ed. J. F. A. Ajayi and M. Crowder. New York: Columbia University Press. Levy, T. E. 2009. The new biblical archaeology: From text to turf. London: Equinox. Lewin, R. 1984. Human evolution: An illustrated introduction. San Francisco: Freeman. Lewin, R. 1988. In the age of mankind. Washington, DC: Smithsonian Institution Press. Lewin, R. 1998. Principles of human evolution. Malden, MA: Blackwell. Lewis-Williams, D. 2002. The mind in the cave: Consciousness and the origins of art. London: Thames & Hudson. Lieberman, D. E. 2001. Another face in our family tree. Nature 410:419–420. Lieberman, P. 1991. Uniquely human: The evolution of speech, thought, and selfless behavior. Cambridge, MA: Harvard University Press. Lipe, W. 1983. The Southwest. In Ancient North Americans, ed. J. Jennings. San Francisco: Freeman. Lister, R. H., and F. C. Lister. 1981. Chaco Canyon, archaeology and archaeologists. References
R-13
_
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
Albuquerque: University of New Mexico Press. Little, B. J. 2002. Public benefits of archaeology. Gainesville: University Press of Florida. Liu, L. 2004. The Chinese Neolithic. Cambridge: Cambridge University Press. Lloyd, S., and F. Safar. 1943. Tell Uqair: Excavations by the Iraq government directorate of antiquities in 1940 and 1941. Journal of Near Eastern Studies 2:131–189. Loewe, M., and E. L. Schaughnessy. 1999. The Cambridge history of ancient China. Cambridge: Cambridge University Press. Londo, J. P., Y. C. Chiang, K. H. Hung, T. Y. Chiang, and B. A. Schaal. 2006. Phylogeography of Asian wild rice, Oriyza rufipogon, reveals multiple independent domestications of cultivated rice, Oryza sativa. Proceedings of the National Academy of Sciences 103(25):9578–9583. Long, A., B. F. Benz, D. J. Donahue, A. J. T. Jull, and L. J. Toolin. 1989. First direct AMS dates on early maize from Tehuacán, Mexico. Radiocarbon 31:1035–1040. Lovejoy, C. O. 1981. The origin of man. Science 211:341–350. Lu, H., X. Yang, M. Ye, K. Liu, Z. Xia, X. Ren, L. Cai, N. Wu, and T. Liu. 2005. Millet noodles in late Neolithic China. Nature 437:967–968. Lumbreras, L. 1974. The peoples and cultures of ancient Peru, trans. B. J. Meggers. Washington, DC: Smithsonian Institution Press. Lynch, T. F. 1980. Guitarrero Cave: Early man in the Andes. London: Academic Press. Lynch, T. F., R. Gillespie, J. A. J. Gowlett, and R. E. M. Hedges. 1985. Chronology of Guitarrero Cave, Peru. Science 229:864–867. MacNeish, R. S. 1978. The science of archaeology? North Scituate, MA: Duxbury Press. MacNeish, R. S. 1981. Tehuacán’s accomplishments. In Supplement to the handbook of Middle American Indians, vol. 1, ed. J. A. Sabloff. Austin: University of Texas Press. MacNeish, R. S., and J. G. Libby, eds. 1995. Origins of rice agriculture: The preliminary report of the Sino-American Jiangxi (PRC) Project. Publications in Anthropology No. 13. El Paso: Centennial Museum, University of Texas.
R-14
References
Page R-14
MacNeish, R. S., F. A. Peterson, and K. V. Flannery. 1970. Prehistory of the Tehuacán Valley. Vol. 3, Ceramics, ed. R. S. MacNeish. Austin: University of Texas Press. Majewski T., and D. Gaimster, eds. 2009. International handbook of historical archaeology. New York: Springer. Malek, J., ed. 1993. Cradles of civilization: Egypt. Norman: University of Oklahoma Press. Maloney, B. K., C. F. W. Higham, and R. Bannanurag. 1989. Early rice cultivation in Southeast Asia: Archaeological and palynological evidence from the Bang Pakong Valley, Thailand. Antiquity 63:363–370. Mann, C. C. 2003. Cracking the khipu code. Science 300:1650–1651. Mann, C. C. 2005. 1491: New revelations of the Americas before Columbus. New York: Knopf. Mannion, A. M. 1999. Domestication and the origins of agriculture: An appraisal. Progress in Physical Geography 23:37–56. Manzanilla, L. 1997. Corporate groups and domestic activities at Teotihuacan. Latin American Antiquity 7:228–246. Manzanilla, L., L. Barba, R. Chávez, A. Tejero, G. Cifuentes, and N. Peralta. 1994. Caves and geophysics: An approximation of the underworld of Teotihuacan, Mexico. Archaeometry 36:141–157. Marcus, J. 1976a. Emblem and state in the Classic Maya Lowlands: An epigraphic approach to territorial organization. Washington, DC: Dumbarton Oaks. Marcus, J. 1976b. The origins of Mesoamerican writing. Annual Review of Anthropology 5:35–67. Marcus, J. 1980. Zapotec writing. Scientific American 242(2):50–64. Marcus, J. 1983a. The conquest slabs of Building J, Monte Albán. In The cloud people: Divergent evolution of the Mixtec and Zapotec civilizations, ed. K. V. Flannery and J. Marcus. New York: Academic Press. Marcus, J. 1983b. Lowland Maya archaeology at the crossroads. American Antiquity 48:454–488. Marcus, J. 1987. Prehistoric fishermen in the kingdom of Huarco. American Scientist 75:393–401. Marcus, J. 1989. Zapotec chiefdoms and the nature of Formative religions. In Regional perspectives on the Olmec, ed. R. J. Sharer and D. C. Grove.
Cambridge: Cambridge University Press. Marcus, J. 1992a. Mesoamerican writing systems: Propaganda, myth, and history in four ancient civilizations. Princeton, NJ: Princeton University Press. Marcus, J. 1992b. Political fluctuations in Mesoamerica. National Geographic Research and Exploration 8:392–411. Marcus, J., ed. 1990. Debating Oaxaca archaeology. Anthropological Papers of the Museum of Anthropology, No. 84. Ann Arbor: University of Michigan. Marcus, J. 2008. The archaeological evidence for social evolution. Annual Review of Anthropology 37:251–266. Marcus, J., and K. V. Flannery. 1996. Zapotec civilization: How urban society evolved in Mexico’s Oaxaca Valley. London: Thames & Hudson. Marcus, J., and J. A. Sabloff, eds. 2008. The ancient city: New perspectives on urbanism in the Old and New Worlds. Santa Fe, NM: School for Advanced Research Press. Marean, C. W., M. Bar-Matthews, J. Bernatchez, E. Fisher, P. Goldberg, A. I. R. Herries, Z. Jacobs, A. Jerardino, P. Karkanas, T. Minichillo, P. J. Nilssen, E. Thompson, I. Watts, and H. M. Williams. 2007. Early human use of marine resources and pigment in South Africa during the Middle Pleistocene. Nature 449:908–910. Marshack, A. 1972a. The roots of civilization. New York: McGraw-Hill. Marshack, A. 1972b. Upper Paleolithic symbol and notation. Science 178:817–828. Martin, P. S., and F. Plog. 1973. The archaeology of Arizona: A study of the Southwest region. New York: Natural History Press. Martin, P. W., and H. E. Wright Jr. 1967. Pleistocene extinctions: The search for a cause. New Haven, CT: Yale University Press. Martin, S., and N. Grube. 1995. Maya superstates. Archaeology 48(6):41–46. Marx, K. 1963. The eighteenth Brumaire of Louis Bonaparte. New York: International Publishers. Mason, R. J. 1981. Great Lakes archaeology. New York: Academic Press. Masuda, S., I. Shimada, and C. Morris. 1985. Andean ecology and civilization: An interdisciplinary perspective on Andean ecological complementarity. Tokyo: University of Tokyo Press. Matheny, R. T., ed. 1980. El Mirador, Petén, Guatemala: An interim report.
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
Papers 45. Provo, UT: New World Archaeological Foundation. Matheny, R. T. 1986. Investigations at El Mirador, Petén, Guatemala. National Geographic Research 2:332–353. Mathien, F. J., and R. H. McGuire, eds. 1986. Ripples in the Chichimec Sea: New considerations of SouthwesternMesoamerican interactions. Carbondale: Southern Illinois University Press. Matos Moctezuma, E. 1984. The great temple of Tenochtitlán. Scientific American 251(2):80–89. Matsuoka, Y., Y. Vigouroux, M. Goodman, J. Sanchez, G. E. Buckler, and J. Doebley. 2002. A single domestication for maize shown by multilocus microsatellite genotyping. Proceedings of the National Academy of Sciences 99(9):6080–6084. Mayr, E. 1970. Population, species, and evolution. Cambridge, MA: Harvard University Press. McDonald, K. 1996. Early Iron-Age settlement of sub-Saharan Africa. In The Oxford companion to archaeology, ed. B. M. Fagan. New York: Oxford University Press. McHenry, H. M. 1982. The pattern of human evolution: Studies on bipedalism, mastication, and encephalization. Annual Review of Anthropology 11:151–173. McIntosh, S. K., ed. 1995. Excavations at Jenné-jeno, Hambarketolo, and Kaniana (Inland Niger Delta, Mali), the 1981 seasons. Berkeley: University of California Press. McIntosh, S. K., ed. 1999. Beyond chiefdoms: Pathways to complexity in Africa. Cambridge: Cambridge University Press. McIntosh, S. K., and R. J. McIntosh. 1980. Jenné-jeno: An ancient African city. Archaeology 33(1):8–14. McIntosh, S. K., and R. J. McIntosh. 1981. West African prehistory. American Scientist 69:602–612. McIntosh, S. K., and R. J. McIntosh. 1983. Current directions in West African prehistory. Annual Review of Anthropology 12:215–258. McIntosh, S. K., and R. J. McIntosh. 1984. The early city in West Africa: Towards an understanding. African Archaeological Review 2:73–98. McIntosh, S. K., and R. J. McIntosh. 1993. Cities without citadels: Understanding urban origins along the middle Niger. In The archaeology of
Page R-15
Africa: Food, metals and towns, ed. T. Shaw, P. Sinclair, B. Andah, and A. Okpoko. London: Routledge. Mckillop, H. 2004. The ancient Maya. Santa Barbara, CA: ABC-Clio. McManamon, F. P., L. S. Cordell, K. G. Lightfoot, and G. R. Milner, eds. 2008. Archaeology in America: An encyclopedia. Westport, CT: Greenwood. Mead, J. I., and D. J. Meltzer. 1985. Environments and extinctions: Man in late glacial North America. Orono, ME: Center for the Study of Early Man. Meadow, R. H. 1984. Animal domestication in the Middle East: A view from the eastern margin. In Animals and archaeology 3, ed. J. Clutton-Brock and C. Grigson. Oxford: British Archaeological Reports. Meggers, B. J., and C. Evans. 1957. Archaeological investigations at the mouth of the Amazon. Bureau of American Ethnology, Bulletin 167. Washington, DC: Smithsonian Institution. Mellaart, J. 1964. Excavations at Çatal Hüyük, 1963: Third preliminary report. Anatolian Studies 14:39–119. Mellaart, J. 1967. Çatal Hüyük: A Neolithic town in Anatolia. London: Thames & Hudson. Mellaart, J. 1975. The Neolithic of the Near East. New York: Scribner. Mellars, P. 2006. A new radiocarbon revolution and the dispersal of modern humans in Eurasia. Nature 439:931–935. Mellars, P., and C. Stringer. 1989. The human revolution: Behavioural and biological perspectives on the origins of modern humans. Edinburgh: Edinburgh University Press. Meltzer, D. J. 2006. Folsom. Berkeley: University of California Press. Meltzer, D. J. 2009. First peoples in a New World: Colonizing Ice Age America. Berkeley: University of California Press. Mendelssohn, K. 1974. The riddle of the pyramids. New York: Praeger. Menotti, F. 2004. Living on the lake in prehistoric Europe. London: Routledge. Mercader, J., H. Barton, J. Gillespie, J. Harris, S. Kuhn, R. Tyler, and C. Boesch. 2007. 4,300-year-old chimpanzee sites and the origins of percussive stone technology. Proceedings of the National Academy of Science 104:3043–3048. Merriman, N., and T. Schadla-Hall. 2002. Public archaeology. London: Routledge.
Milisauskas, S. 1978. European prehistory. New York: Academic Press. Milisauskas, S., ed. 2002. European prehistory, a survey. New York: Springer. Millon, R. 1967. Teotihuacan. Scientific American 216(6):38–48. Millon, R. 1973. Urbanization at Teotihuacan, Mexico. Vol. 1, The Teotihuacan map. Austin: University of Texas Press. Millon, R. 1976. Social relations in ancient Teotihuacan. In The Valley of Mexico, ed. E. R. Wolf. Albuquerque: University of New Mexico Press. Millon, R. 1981. Teotihuacan: City, state, and civilization. In Supplement to the handbook of Middle American Indians. Vol. 1, Archaeology, ed. J. A. Sabloff. Austin: University of Texas Press. Mills, B. J. 2002. Recent research on Chaco: Changing views on economy, ritual, and society. Journal of Archaeological Research 10:65–117. Milner, G. R. 1998. The Cahokia chiefdom: The archaeology of a Mississippian society. Washington, DC: Smithsonian Institution Press. Milner, G. R. 2003. Archaeological indicators of rank in the Cahokia chiefdom. In Theory, method, and practice in modern archaeology, ed. R. J. Jeske and D. K. Charles. Westport, CT: Praeger. Milner, G. R. 2004. The moundbuilders: Ancient peoples of eastern North America. London: Thames & Hudson. Montague, A. 1964. The concept of race. New York: Free Press. Moore, A. M. T. 1985. The development of Neolithic societies in the Near East. Advances in World Archaeology 4:1–70. Moore, A. M. T., G. C. Hillman, and A. J. Legge. 2000. Village on the Euphrates: The excavation of Abu Hureyra. Oxford: Oxford University Press. Moore, C. B. 1905. Certain aboriginal remains of the Black Warrior River. Journal of the Academy of Natural Sciences of Philadelphia 13:125–244. Moore, J. D. 1996. Architecture and power in the ancient Andes: The archaeology of public buildings. Cambridge: Cambridge University Press. Moorehead, W. K. 1922. The Hopewell mound group of Ohio. Anthropological Series 6(5). Publication 211. Chicago: Field Museum of Natural History. Reprinted in 1968. Morley, S. G. (ed. R. J. Sharer). 1994. The ancient Maya, 5th ed. Stanford, CA: Stanford University Press. References
R-15
_
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
Morley, S. G., and G. W. Brainerd. 1956. The ancient Maya, 3d ed. Stanford, CA: Stanford University Press. Morrell, V. 2001. The pyramid builders. National Geographic 200(5):78–99. Morris, C. 1998. Inka strategies of incorporation and governance. In Archaic states, ed. G. M. Feinman and J. Marcus. Santa Fe, NM: School of American Research Press. Morris, C., and D. E. Thompson. 1985. Huánuco Pampa: An Inca city and its hinterland. London: Thames & Hudson. Morris, C., and A. von Hagen. 1993. The Inka empire and its Andean origins. New York: Abbeville Press. Moseley, M. E. 1975a. Chan Chan: Andean alternative of the preindustrial city? Science 187:219–225. Moseley, M. E. 1975b. The maritime foundations of Andean civilization. Menlo Park, CA: Benjamin/Cummings. Moseley, M. E. 1975c. Prehistoric principles of labor organization in the Moche Valley, Peru. American Antiquity 40:191–196. Moseley, M. E. 1983. Central Andean civilization. In Ancient South Americans, ed. J. D. Jennings. San Francisco: Freeman. Moseley, M. E. 2001. The Incas and their ancestors: The archaeology of Peru, rev. ed. London: Thames & Hudson. Moseley, M. E., and K. C. Day, eds. 1982. Chan Chan: Andean desert city. Albuquerque: University of New Mexico Press. Movius, H. L. 1948. The Lower Paleolithic culture of southern and eastern Asia. Transactions of the American Philosophical Society 38:329–351. Muller, J. 1983. The Southeast. In Ancient North Americans, ed. J. Jennings. San Francisco: Freeman. Mulvaney, D. J. 1975. The prehistory of Australia, 2d ed. Baltimore: Pelican. Mulvaney, J., and J. Kamminga. 1999. The prehistory of Australia. Washington, DC: Smithsonian Institution Press. Muro, M. 1998. New finds explode old views of the American Southwest. Science 279:653–654. Murowchick, R. E., ed. 1994. Cradles of civilization: China. Norman: University of Oklahoma Press. Murra, J. V. 1962. Cloth and its function in the Inca state. American Anthropologist 64:710–728.
R-16
References
Page R-16
Murra, J. V. 1972. El “control vertical” de un máximo de pisos ecológicos en la economía de las sociedades andinas. In Visita de la provincia de Leon de Huanuco (1562), vol. 2, ed. J. V. Murra. Huánuco, Peru: Universidad Nacional Hermilio Valdizan. Ndoro, W. 1996. Great Zimbabwe. In The Oxford companion to archaeology, ed. B. M. Fagan. New York: Oxford University Press. Neitzel, J. 1989. The Chacoan regional system: Interpreting the evidence for social complexity. In The sociopolitical structure of prehistoric southwestern societies, ed. S. Upham, K. G. Lightfoot, and R. A. Jewett. Boulder, CO: Westview. Neitzel, J. E., ed. 1999. Great towns and regional polities in the prehistoric American Southwest and Southeast. Albuquerque: University of New Mexico Press. Neuman, R. W. 1990. An introduction to Louisiana archaeology. Baton Rouge: Louisiana State University Press. Neumann, T. N., and R. M. Sanford. 2001. Cultural resources archaeology. Walnut Creek, CA: AltaMira Press. Nicholas, L. M., and G. M. Feinman. 1989. A regional perspective on Hohokam irrigation in the lower Salt River Valley, Arizona. In The sociopolitical structure of prehistoric southwestern societies, ed. S. Upham, K. G. Lightfoot, and R. A. Jewett. Boulder, CO: Westview. Nissen, H. J. 1986. The archaic texts from Uruk. World Archaeology 17:317–334. Nissen, H. J. 1988. The early history of the ancient Near East, 9000–2000 B.C. Chicago: University of Chicago Press. Noble, D. G., ed. 2004. In search of Chaco: New approaches to an archaeological enigma. Santa Fe, NM: School for American Research Press. Normile, D. 1997. Yangtze seen as earliest rice site. Science 275:309. Oakley, K. P. 1955. Fire as a Paleolithic tool and weapon. Proceedings of the Prehistoric Society 21:36–48. Oates, J. 1980. The emergence of cities in the Near East. In The Cambridge encyclopedia of archaeology, ed. A. Sherratt. New York: Crown. Oates, J., A. McMahon, P. Karsgaard, S. Al Quntar, and J. Ur. 2007. Early Mesopotamian urbanism: A new view from the north. Antiquity 81:585–600.
O’Connor, D. 1980. Egypt and the Levant in the Bronze Age. In The Cambridge encyclopedia of archaeology, ed. A. Sherratt. New York: Crown. Oliver, R., and B. M. Fagan. 1975. Africa in the Iron Age: c. 500 B.C. to A.D. 1400. Cambridge: Cambridge University Press. Ortiz de Montellano, B. 1978. Aztec cannibalism: An ecological necessity? Science 200:611–617. Osborne, R. 2009. Greece in the making 1200–479 B.C. London: Routledge. Otto, M. P. 1979. Hopewell antecedents in the Adena heartland. In Hopewell archaeology: The Chillicothe conference, ed. D. S. Brose and N. Greber. Kent, OH: Kent State University Press. Parkington, J. E. 1972. Seasonal mobility in the Late Stone Age. African Studies 31:223–243. Parkington, J. E. 1981. Stone tools and resources: A case study from South Africa. World Archaeology 13:16–30. Parkington, J. E. 1984. Changing views of the Later Stone Age of South Africa. Advances in World Archaeology 3:89–142. Parpola, A. 1986. The Indus script: A challenging puzzle. World Archaeology 17:399–419. Parsons, J. R. 1972. Archaeological settlement patterns. Annual Review of Anthropology 1:127–150. Pauketat, T. R. 1998. Refiguring the archaeology of greater Cahokia. Journal of Archaeological Research 6:45–89. Pauketat, T. R. 2003. Resettled farmers and the making of a Mississippian polity. American Antiquity 68:39–66. Pauketat, T. R. 2004a. Ancient Cahokia and the Mississippians. Cambridge: Cambridge University Press. Pauketat, T. R. 2004b. The economy of the moment: Cultural practices and Mississippian chiefdoms. In Archaeological perspectives on political economies, ed. G. M. Feinman and L. M. Nicholas. Salt Lake City: University of Utah Press. Pauketat, T. R. 2009. Cahokia: Ancient America’s great city on the Mississippi. New York: Viking Adult. Pauketat, T. R., and T. E. Emerson, eds. 1997. Cahokia: Domination and ideology in the Mississippian world. Lincoln: University of Nebraska Press. Pauketat, T. R., and N. H. Lopinot. 1997. Cahokian population dynamics. In Cahokia: Domination and ideol-
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
ogy in the Mississippian world, ed. T. R. Pauketat and T. E. Emerson. Lincoln: University of Nebraska Press. Paul, A., and S. A. Turpin. 1986. The ecstatic shaman theme of Paracas textiles. Archaeology 39(5):20–27. Pearsall, D. 1989. Paleoethnobotany. Orlando, FL: Academic Press. Pearsall, D. 1992. The origins of plant cultivation in South America. In Origins of agriculture in world perspective, ed. C. W. Cowan and P. J. Watson. Washington, DC: Smithsonian Institution Press. Pearson, M. P. 1996. Mortuary analysis. In The Oxford companion to archaeology, ed. B. M. Fagan. New York: Oxford University Press. Pearson, M. P., R. Cleal, P. Marshall, S. Needham, J. Pollard, C. Richards, C. Ruggles, A. Sheridan, J. Thomas, C. Tilley, K. Welham, A. Chamberlain, C. Chenery, J. Evans, and C. Knüsel. 2007. The age of Stonehenge. Antiquity 811(313):617–639. Pearson, R., and A. Underhill. 1987. The Chinese Neolithic: Recent trends in research. American Anthropologist 89:807–822. Peebles, C. S., and C. A. Black. 1987. Moundville from 1000–1500 A.D. as seen from 1840 to 1985 A.D. In Chiefdoms in the Americas, ed. R. D. Drennan and C. A. Uribe. Lanham, MD: University Press of America. Peebles, C. S., and S. Kus. 1977. Some archaeological correlates of ranked society. American Antiquity 42:421–448. Peltenberg, E., S. Colledge, P. Croft, A. Jackson, C. McCartney, and M. A. Murra. 2000. Agro-pastoralist colonization of Cyprus in the 10th millennium B.P.: Initial assessments. Antiquity 74:844–853. Peregrine, P. N., and M. Ember, eds. 2003. Encyclopedia of prehistory. New York: Springer Verlag. Perlès, C. 2001. The Early Neolithic in Greece. Cambridge: Cambridge University Press. Perony, D. 1930. Le Moustier. Revue Anthropologique 14. Perrot, J. 1966. Le gisement natoufien de Mallaha (Eynan), Israel. L’Anthropologie 47:437–484. Perry, L., R. Dickau, S. Zarrillo, I. Holst, D. M. Pearsall, D. R. Piperno, M. J. Berman, R. G. Cooke, K. Rademaker, A. J. Ranere, J. S. Raymond, D. H. Sandweiss, F. Scaramelli, K. Tarble,
Page R-17
and J. A. Zeidler. 2007. Starch fossils and the domestication and dispersal of chili peppers (Capsicum sp. L) in the Americas. Science 315:986–988. Perry, L., D. H. Sandweiss, D. R. Piperno, K. Rademaker, M. A. Malpass, A. Umire, and P. de la Vera. 2006. Early maize agriculture and interzonal interaction in southern Peru. Nature 440:76–79. Peterson, I. 1988. Tokens of plenty. Science News 134:408–410. Pfeiffer, J. E. 1982. The creative explosion: An enquiry into the origins of art and religion. New York: Harper & Row. Pfeiffer, J. E. 1985. The emergence of humankind. New York: Harper & Row. Phillipson, D. 1977. The later prehistory of eastern and southern Africa. London: Heinemann. Phillipson, D. 1980. Iron Age Africa and the expansion of the Bantu. In The Cambridge encyclopedia of archaeology, ed. A. Sherratt. New York: Crown. Phillipson, D. 1985. African archaeology. Cambridge: Cambridge University Press. Phillipson, D. W. 1996. Prehistory of Africa. In The Oxford companion to archaeology, ed. B. M. Fagan. New York: Oxford University Press. Pickering, T. R., T. D. White, and N. Toth. 2000. Cutmarks on a Plio-Pleistocene hominid from Sterkfontein, South Africa. American Journal of Physical Anthropology 111:579–584. Pickford, M., and B. Senut. 2001. The geological and faunal context of Late Miocene hominid remains from Lukeino, Kenya. Comptes Rendus de l’Académie de Sciences 332:145–152. Pikirayi, I. 2006. The demise of Great Zimbabwe, A.D. 1420–1550: An environmental re-appraisal. In Cities in the world, 1500–2000, ed. A. Green and R. Leech. Leeds, UK: Maney. Pilbeam, D. 1985. Distinguished lecture: Hominoid evolution and hominoid origins. American Anthropologist 88:295–312. Piperno, D. R., and K. V. Flannery. 2001. The earliest archaeological maize (Zea mays L.) from highland Mexico: New accelerator mass spectrometry dates and their implications. Proceedings of the National Academy of Sciences 98:2101–2103. Pitts, M. 2008. Stonehenge: One of our largest excavations draws to a close. British Archaeology. York, England: Council for British Archaeology.
Pleger, T. C. 2000. Old copper and red ocher social complexity. Midcontinental Journal of Archaeology 25:169–190. Plog, S. 2008. Ancient peoples of the American Southwest, 2d ed. London: Thames & Hudson. Pollock, S. 1999. Ancient Mesopotamia: The Eden that never was. Cambridge: Cambridge University Press. Pope, K. O., M. E. D. Pohl, J. G. Jones, D. L. Lentz, C. von Nagy, F. J. Vega, and I. R. Quitmyer. 2001. Origin and environmental setting of ancient agriculture in the lowlands of Mesoamerica. Science 292:1370–1373. Popson, C. P. 2002. Grim rites of the Moche. Archaeology 55(2):30–35. Portal, J., ed. 2007. The first emperor: China’s terracotta army. London: British Museum Press. Possehl, G. L. 1990. Revolution in the urban revolution: The emergence of Indus urbanization. Annual Review of Anthropology 19:261–282. Possehl, G. L. 1997. The transformation of the Indus civilization. Journal of Archaeological Research 11:425–472. Possehl, G. L. 2002. The Indus civilization: A contemporary perspective. Walnut Creek, CA: AltaMira. Postgate, J. N. 1994. Early Mesopotamia: Society and economy at the dawn of history. New York: Routledge. Powell, T. G. E. 1980. The Celts. London: Thames & Hudson. Praetzellis, A. 2000. Death by theory: A tale of mystery and archaeological theory. Walnut Creek, CA: AltaMira Press. Prag, J., and R. Neeve. 1999. Making faces. London: British Museum Press. Price, T. D. 1987. The Mesolithic of western Europe. Journal of World Prehistory 1:225–305. Price, T. D., ed. 1989. The chemistry of prehistoric bone. Cambridge: Cambridge University Press. Price, T. D., ed. 2000. Europe’s first farmers. Cambridge: Cambridge University Press. Price, T. D., and E. Brinch Petersen. 1987. A Mesolithic community in Denmark. Scientific American 255(3): 111–121. Price, T. D., and J. A. Brown, eds. 1985. Prehistoric hunter-gatherers. Orlando, FL: Academic Press. Price, T. D., and A. B. Gebauer. 2005. Smakkerup Huse: A coastal Late Mesolithic site in Denmark. Aarhus: Aarhus University Press. References
R-17
_
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
Price, T. D., and A. B. Gebauer, eds. 1995. Last hunters—first farmers: New perspectives on the prehistoric transition to agriculture. Santa Fe, NM: School for American Research Press. Pringle, H. 1997. Oldest mound complex found at Louisiana site. Science 277:1761–1762. Prufer, K. M., and J. E. Brady. 2005. Stone houses and earth lords: Maya religion in the cave context. Boulder: University Press of Colorado. Prufer, O. 1964. The Hopewell cult. Scientific American 211(6):90–102. Puleston, D. E. 1973. Ancient Maya settlement patterns and environment at Tikal, Guatemala: Implications for subsistence models. Ph.D. diss., University of Pennsylvania, Philadelphia. Puleston, D. E. 1977. The art and technology of hydraulic agriculture in the Maya Lowlands. In Social processes and Maya prehistory, ed. N. Hammond. New York: Academic Press. Puleston, D. E. 1978. Terracing, raised fields, and tree cropping in the Maya Lowlands: A new perspective on the geography of power. In Prehispanic Maya agriculture, ed. P. D. Harrison and B. L. Turner II. Albuquerque: University of New Mexico Press. Quilter, J. 1985. Architecture and chronology at El Paraíso, Peru. Journal of Field Archaeology 12:279–297. Quilter, J. 1991. Late preceramic Peru. Journal of World Prehistory 5:387–438. Quilter, J., B. Ojeda, D. M. Pearsall, D. H. Sandweiss, J. G. Jones, and E. S. Wing. 1991. Subsistence economy of El Paraíso, an early Peruvian site. Science 251:277–283. Quilter, J., and T. Stocker. 1983. Subsistence economies and the origins of agriculture. American Anthropologist 85:545–562. Rambo, A. T. 1991. The study of cultural evolution. In Profiles in cultural evolution: Papers from a conference in honor of Elman R. Service, ed. A. T. Rambo and K. Gillogly. Anthropological Papers of the Museum of Anthropology, No. 85. Ann Arbor: University of Michigan. Randsborg, K. 1975. Social dimensions of early Neolithic Denmark. Proceedings of the Prehistoric Society 41:105–118. Rawson, J. 2008. The first emperor’s tomb: The afterlife universe. In The
R-18
References
Page R-18
first emperor: China’s terracotta army, ed. J. Portal. London: British Museum Press. Raymond, J. S. 1981. The maritime foundations of Andean civilization: A reconsideration of the evidence. American Antiquity 46:806–821. Redman, C. 1978. The rise of civilization: From early farmers to urban society in the ancient Near East. San Francisco: Freeman. Redmond, E. M. 1983. A fuego y sangre: Early Zapotec imperialism in the Cuicatlán Cañada, Oaxaca. Memoirs of the Museum of Anthropology, No. 16. Ann Arbor: University of Michigan. Reed, C. A., ed. 1977. The origins of agriculture. The Hague: Mouton. Renfrew, C. 1974. Before civilization. New York: Knopf. Renfrew, C., and P. Bahn. 1998. Archaeology: Theories, methods, and practice. London: Thames & Hudson. Renfrew, J. 1973. Palaeoethnobotany. New York: Columbia University Press. Renfrew, C. 2009. Prehistory: The making of the human mind. Modern Library Chronicles. Rice, G. 1987. La Ciudad: A perspective on Hohokam community systems. In The Hohokam village: Site structure and organization, ed. D. E. Doyel. Glenwood Springs, CO: Southwestern and Rocky Mountain Division of the American Association for the Advancement of Science. Richards, J. C. 2007. Stonehenge: The story so far. Swindon, England: English Heritage. Richards, M. P., P. B. Pettitt, M. C. Stiner, and E. Trinkaus. 2001. Stable isotope evidence for increasing dietary breadth in the European mid–Upper Paleolithic. Proceedings of the National Academy of Science 98:6528–6532. Richerson, P. J., and R. Boyd. 2001. Institutional evolution in the Holocene: The rise of complex societies. In The origin of human social institutions, ed. W. G. Runciman. Oxford: Oxford University Press. Rindos, D. 1984. The origins of agriculture: An evolutionary perspective. New York: Academic Press. Robson, J. R. K., R. I. Ford, K. V. Flannery, and J. E. Konlande. 1976. The nutritional significance of maize and teosinte. Ecology of Food and Nutrition 4:243–249.
Roe, D. 1981. The Lower and Middle Paleolithic periods in Britain. London: Routledge. Rogers, A. R., D. Iltis, and S. Wooding. 2004. Genetic variation at the MC1R locus and the time since loss of human body hair. Current Anthropology 45:105–108. Rolland, N., and H. L. Dibble. 1990. A new synthesis of Middle Paleolithic variability. American Antiquity 55:480–499. Rollefson, G. O. 1985. The 1983 season at the Early Neolithic site of ’Ain Ghazal. National Geographic Research 1(1):44–62. Rollo, F., M. Ubaldi, L. Ermini, and I. Marota. 2002. Ötzi’s last meals: DNA analysis of the intestinal content of the Neolithic glacier mummy from the Alps. Proceedings of the National Academy of Sciences 99:12594–12599. Ronen, A., ed. 1982. The transition from Lower to Middle Paleolithic and the origin of modern man. Oxford: British Archaeological Reports. Roosevelt, A. 1989. Lost civilizations of the lower Amazon. Natural History 98(2):74–82. Roosevelt, A. C., R. A. Housley, M. Imazio da Silveira, S. Maranca, and R. Johnson. 1991. Eighth millennium pottery from a prehistoric shell midden in the Brazilian Amazon. Science 254:1621–1624. Rosenberg, M., R. Nesbitt, R. W. Redding, and B. L. Peasnall. 1998. Hallan Çemi, pig husbandry and post-Pleistocene adaptations along the Taurus-Zagros archaeology (Turkey). Paléorient 24:25–41. Rosman, A., and P. G. Rubel. 1971. Feasting with mine enemy: Rank and exchange among Northwest Coast societies. New York: Columbia University Press. Rothman, M. S., ed. 2001. Uruk Mesopotamia and its neighbors: Crosscultural interactions in the era of state formation. Santa Fe, NM: School of American Research Press. Rouse, I. 1992. The Tainos: Rise and decline of the people who greeted Columbus. New Haven, CT: Yale University Press. Rowe, J. H. 1947. Inca culture at the time of Spanish conquest. In Handbook of South American Indians. Vol. 2, The Andean civilizations, ed.
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
J. H. Steward. Bureau of American Ethnology, Bulletin 143. Washington, DC: Smithsonian Institution. Rowe, J. H. 1967. What kind of settlement was Inca Cuzco? Nawpa Pacha 5:59–76. Rowe, J. H. 1987. Machu Pijchu: A la luz de los documentos del siglo XVI. Kuntur 4:12–20. Ruddiman, W. F., and J. E. Kutzbach. 1991. Plateau uplift and climatic change. Scientific American 264(3):66–75. Ruz Lhuillier, A. 1973. El Templo de las Inscripciones: Palenque. Mexico City: Instituto Nacional de Antropología e Historia. Sabloff, J. A. 1990. The new archaeology and the ancient Maya. New York: Freeman. Sabloff, J. A., ed. 2003. Tikal: Dynasties, foreigners, and affairs of state. Santa Fe, NM: School of American Research Press. Sagan, C. 1987. Billions and billions. Parade, May 31, 9. Sage, R. F. 1995. Was low atmospheric CO2 during the Pleistocene a limiting factor for the origin of agriculture? Global Change Biology 1:93–106. Sahagún, F. B. 1950–1982. Florentine codex: General history of the things of New Spain, trans. A. J. O. Anderson and C. E. Dibble. 11 vols. Santa Fe, NM: School for American Research; Provo: University of Utah Press. Sahlins, M. D. 1968. Notes on the original affluent society. In Man the hunter, ed. R. B. Lee and I. DeVore. Chicago: Aldine. Sahlins, M. D. 1972. Stone Age economics. Chicago: Aldine. Sanders, W. T., J. R. Parsons, and R. S. Santley. 1979. The Basin of Mexico: Ecological processes in the evolution of a civilization. New York: Academic Press. Sanders, W. T., and B. Price. 1968. Mesoamerica: The evolution of a civilization. New York: Random House. Sarich, V. 1983. Retrospective on hominid macromolecular systematics. In New interpretations of ape and human ancestry, ed. R. L. Ciochon and R. S. Corrucini. New York: Plenum. Sassaman, K. 2004. Complex huntergatherers in evolution and history: A North American perspective. Journal of Archaeological Research 12:227–280. Sassaman, K. 2005. Poverty Point as structure, event, process. Journal of
Page R-19
Archaeological Method and Theory 12:335–364. Saturno, W. A., K. A. Taube, D. Stuart, and H. Hurst. 2005. The murals of San Bartolo, El Petén, Guatemala. Part 1, The north wall. Ancient America, No. 7. Barnardsville, NC: Center for Ancient American Studies. Sauer, C. O. 1952. Agricultural origins and dispersals. New York: American Geographical Society. Saunders, J. W., and T. Allen. 1994. Hedgepeth mounds, an Archaic mound complex in north-central Louisiana. American Antiquity 59:471–489. Saunders, J. W., R. D. Mandel, R. T. Saucier, E. T. Allen, C. T. Hallmark, J. K. Johnson, E. H. Jackson, C. M. Allen, G. L. Stringer, D. S. Frink, J. K. Feathers, S. Williams, K. J. Gremillion, M. F. Vidrine, and R. Jones. 1997. A mound complex in Louisiana at 5400–5000 years before the present. Science 277:1796–1799. Scarborough, V. L., and D. R. Wilcox, eds. 1991. The Mesoamerican ballgame. Tucson: University of Arizona Press. Scarre, C., and G. Scarre. 2006. The ethics of archaeology. Cambridge: Cambridge University Press. Schele, L., and M. E. Miller. 1986. The blood of kings: Dynasty and ritual in Maya art. Fort Worth, TX: Kimball Art Museum. Schmandt-Besserat, D. 1978. The earliest precursor of writing. Scientific American 238(6):50–59. Schmandt-Besserat, D. 1980. The envelopes that bear the first writing. Technology and Culture 21:357–385. Schmandt-Besserat, D. 1990. Accounting in the prehistoric Middle East. Archeomaterials 4:15–23. Schrenk, F. 2008. The Neanderthals. London: Routledge. Schwartz, J. H. 1995. Skeleton keys. Oxford: Oxford University Press. Schwartz, J. H., and T. D. White. 2003. Another perspective on hominid diversity. Science 301:763–764. Senut, B., et al. 2001. First hominid from the Miocene (Lukeino Formation, Kenya). C. R. Acad. Sci. Paris, Earth and Planetary Sciences 332:137–144. Serjeantson, D. 2009. Birds. Cambridge Manuals in Archaeology. Cambridge: Cambridge University Press. Service, E. R. 1975. Origins of the state and civilization: The process of cultural evolution. New York: Norton.
Shackleton, N. J., and N. D. Opdyke. 1973. Oxygen isotope and paleomagnetic stratigraphy of equatorial Pacific core V28-238: Oxygen isotope temperatures and ice volume on a 105 and 106 year scale. Quarternary Research 3:39–55. Shady Solis, R., J. Haas, and W. Creamer. 2001. Dating Caral, a preceramic site in the Supe Valley on the central coast of Peru. Science 292:723–726. Sharer, R. J., and D. C. Grove, eds. 1989. Regional perspectives on the Olmec. Cambridge: Cambridge University Press. Sharer, R. J., and L. P. Traxler. 2006. The ancient Maya, 6th ed. Stanford, CA: Stanford University Press. Shea, J. J. 2003. Neandertals, competition, and the origin of modern human behavior in the Levant. Evolutionary Anthropology 12:173–187. Sherratt, A., ed. 1980. The Cambridge encyclopedia of archaeology. New York: Crown. Shipman, P. 1983. Early hominid lifestyle: Hunting and gathering or foraging and scavenging? In Animals and archaeology: Hunters and their prey, ed. J. Clutton-Brock and C. Grigson. Oxford: British Archaeological Reports. Shipman, P. 2002. Hunting the first hominid. American Scientist 90:25–27. Shopland, N. 2006. A finds manual: Excavating, processing and storing. Stroud, UK: Tempus. Shutler, R., Jr. 1983. Early man in the New World. Beverly Hills, CA: Sage Publications. Siemens, A. H., and D. E. Puleston. 1972. Ridged fields and associated features in southern Campeche: New perspectives on the lowland Maya. American Antiquity 37:228–239. Simon, M. 2003. In line of the founder: A view of dynastic politics at Tikal. In Tikal: Dynasties, foreigners, and affairs of state, ed. J. A. Sabloff. Santa Fe, NM: School of American Research Press. Simons, E. 1972. Primate evolution. New York: Macmillan. Simpson, G. G. 1967. The meaning of evolution, rev. ed. New Haven, CT: Yale University Press. Singer, R., and J. Wymer. 1982. The Middle Stone Age at Klasies River Mouth in South Africa. Chicago: University of Chicago Press. Skelton, R. R., H. M. McHenry, and G. M. Drawhorn. 1986. Phylogenetic References
R-19
_
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
analysis of early hominids. Current Anthropology 27:21–43. Smith, B. D. 1986. The archaeology of the southeastern United States: From Dalton to de Soto, 10,500–500 B.P. Advances in World Archaeology 5:1–92. Smith, B. D. 1989. Origins of agriculture in eastern North America. Science 246:1566–1571. Smith, B. D. 1992. Rivers of change: Essays on early agriculture in eastern North America. Washington, DC: Smithsonian Institution Press. Smith, B. D. 1996. Agricultural chiefdoms of the Eastern Woodlands. In The Cambridge history of the native peoples of the Americas. Vol. 1, North America, part 1, ed. B. G. Trigger and W. E. Washburn. Cambridge: Cambridge University Press. Smith, B. D. 1997. Reconsidering the Ocampo caves and the era of incipient cultivation in Mesoamerica. Latin American Antiquity 8:342–383. Smith, B. D. 1998. The emergence of agriculture. New York: Scientific American Library. Smith, B. D. 2001. Documenting plant domestication: The consilience of biological and archaeological approaches. Proceedings of the National Academy of Sciences 98:1324–1326. Smith, B. D. 2005. Reassessing Coxcatlan Cave and the early history of domesticated plants in Mesoamerica. Proceedings of the National Academy of Sciences 02(27):9438–9445. Smith, B. D. 2006. Eastern North America as an independent center of plant domestication. Proceedings of the National Academy of Sciences 103(33):12,223–12,228. Smith, C. E. 1980. Plant remains from Guitarrero Cave. In Guitarrero Cave, ed. T. F. Lynch. New York: Academic Press. Smith, F. H., and F. Spencer, eds. 1984. The origins of modern humans. New York: A. R. Liss. Smith, M. E. 1996. The Aztecs. Malden, MA: Blackwell. Smith, M. E., and M. A. Masson, eds. 2000. The ancient civilizations of Mesoamerica: A reader. Malden, MA: Blackwell. Smith, M. L., ed. 2003. The social construction of ancient cities. Washington, DC: Smithsonian Institution Press.
R-20
References
Page R-20
Snow, D. R. 2009. Archaeology of Native North America. Englewood Cliffs, NJ: Prentice-Hall. Soffer, O. 1985. The Upper Paleolithic of the central Russian plains. New York: Academic Press. Solecki, R. 1971. Shanidar: The first flower people. New York: Knopf. Solheim, W. G., II. 1972a. An earlier agricultural revolution. Scientific American 226(4):34–41. Solheim, W. G., II. 1972b. Early man in Southeast Asia. Expedition 14(3):25–31. Spence, M. W. 1974. Residential practices and the distribution of skeletal traits in Teotihuacan, Mexico. Man 9:262–273. Spence, M. W. 1981. Obsidian production and the state in Teotihuacan. American Antiquity 46:769–787. Spencer, C. S. 1982. The Cuicatlán Cañada and Monte Albán: A study of primary state formation. New York: Academic Press. Spindler, K. 1994. The man in the ice. London: Weidenfeld & Nicholson. Spooner, B., ed. 1972. Population growth: Anthropological implications. Cambridge, MA: MIT Press. Standage, T. 2005. A history of the world in 6 glasses. New York: Walker. Stanford, C. 1998. The social behavior of chimpanzees and bonobos. Current Anthropology 39:399–420. Stanish, C. 2002. Tiwanaku political economy. In Andean archaeology I: Variations in sociopolitical organization, ed. W. H. Isbell and H. Silverman. New York: Kluwer/Plenum. Stanish, C. 2003. Ancient Tiwanaku: The evolution of complex society in southern Peru and northern Bolivia. Berkeley: University of California Press. Stein, G. J. 1998. Heterogeneity, power, and political economy: Some current research issues in the archaeology of Old World complex societies. Journal of Archaeological Research 6:1–44. Stein, G. J. 2001. Understanding ancient state societies in the Old World. In Archaeology at the millennium: A sourcebook, ed. G. M. Feinman and T. D. Price. New York: Kluwer/Plenum. Stein, G. J., and M. S. Rothman. 1994. Chiefdoms and early states in the Near East. Madison, WI: Prehistory Press. Stephens, J. L. 1841. Incidents of travel in Central America, Chiapas, and Yucatán.
2 vols. New York: Harper & Row. Reprint, New York: Dover, 1962. Stephens, J. L. 1843. Incidents of travel in Yucatán. 2 vols. New York: Harper & Row. Reprint, New York: Dover, 1963. Steponaitis, V. 1983. Ceramics, chronology, and community patterns: An archaeological study at Moundville. New York: Academic Press. Steponaitis, V. 1986. Prehistoric archaeology in the southeastern United States, 1970–1985. Annual Review of Anthropology 15:363–404. Steponaitis, V. 1991. Contrasting patterns of Mississippian development. In Chiefdoms: Power, economy, and ideology, ed. T. Earle. Cambridge: Cambridge University Press. Stille, A. 2003. The future of the past. New York: Picador Books. Stirling, M. W. 1943. Stone monuments of southern Mexico. Bureau of American Ethnology, Bulletin 138. Washington, DC: Smithsonian Institution. Stoltman, J. B., ed. 1991. New perspectives on Cahokia: Views from the periphery. Madison, WI: Prehistory Press. Storey, A. A., D. Quiroz, J. M. Ramírez, N. Beavan-Athfield, D. J. Addison, R. Walter, T. Hunt, J. Athens, L. Huynen, and E. A. Matisoo-Smith. 2008. Pre-Columbian chickens, dates, isotopes, and mtDNA. Proceedings of the National Academy of Science 105:E99. Storey, G. R., ed. 2006. Urbanism in the preindustrial world: Cross-cultural approaches. Tuscaloosa: University of Alabama Press. Stringer, C. B. 1985. Middle Pleistocene hominid variability and the origin of Late Pleistocene humans. In Ancestors: The hard evidence, ed. E. Delson. New York: A. R. Liss. Stringer, C. B. 1988. The Neanderthals. London: Thames & Hudson. Stringer, C. B. 1990. The emergence of modern humans. Scientific American 259(12):98–103. Stringer, C., and P. Andrews. 2005. The complete world of human evolution. London: Thames & Hudson. Sudgen, D. E., and B. S. John. 1976. Glaciers and landscape. London: E. Arnold. Susman, R. L., and J. T. Stern. 1982. Functional morphology of Homo habilis. Science 217:931–934. Sutliffe, A. J. 1985. On the track of Ice Age mammals. Cambridge, MA: Harvard University Press.
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
Sutton, M. Q., and B. S. Arkush. 1998. Archaeological laboratory methods. Dubuque, IA: Kendall/Hunt. Swaminathan, M. S. 1984. Rice. Scientific American 250(1):80–93. Tanner, N. 1981. On becoming human. London: Cambridge University Press. Tattersall, I. 1995. The fossil trail: How we know what we think we know about human evolution. Oxford: Oxford University Press. Tattersall, I. 1999. The last Neanderthal: The rise, success, and mysterious extinction of our closest human relatives. Boulder, CO: Westview. Tattersall, I., C. Delson, and J. V. Couvering, eds. 1988. Encyclopedia of human evolution and prehistory. New York: Garland. Tattersall, I. 2001. The human odyssey: Four million years. Lincoln, NE: Universe Press. Tattersall, I., and J. H. Schwartz. 2001. Extinct humans. Boulder, CO: Westview. Tauber, H. 1981. 13C evidence for dietary habits of prehistoric man in Denmark. Nature 292:332–333. Taylor, R. E. 1988. Radiocarbon dating. New York: Academic Press. Taylour, W. 1989. The Mycenaeans. London: Thames & Hudson. Te-k’un, C. 1959. Archaeology in China. Vol. 1, Prehistoric China. Cambridge: W. Heffer & Sons. Te-k’un, C. 1960. Archaeology in China. Vol. 2, Shang burials. Cambridge: W. Heffer & Sons. Te-k’un, C. 1966. Archaeology in China: New light on prehistoric China. Cambridge: W. Heffer & Sons. Tello, J. C. 1943. Discovery of the Chavín culture in Peru. American Antiquity 9:135–160. Teltser, P. 1996. Mississippian culture. In The Oxford companion to archaeology, ed. B. M. Fagan. Oxford: Oxford University Press. Templeton, A. R. 1993. The “Eve” hypothesis: A genetic critique and reanalysis. American Anthropologist 95:51–72. Thiel, J. H., and J. B. Mabry, eds. 2006. Rio Hondo archaeology, 2000–2003: Investigations at the San Agustín Mission and Mission Gardens, Tucson Presidio, Tucson Pressed Brick Company, and Clearwater site. Technical report No. 2004-11. Tucson, AZ: Desert Archaeology.
Page R-21
Thieme, H. 1997. Lower Paleolithic hunting spears from Germany. Nature 385:807. Thomas, D. H. 1983. The archaeology of Monitor Valley 2: Gatecliff Shelter. New York: American Museum of Natural History. Thomas, D. H. 1989. Archaeology. New York: Holt, Rinehart & Winston. Thomas, D. H. 2000. Skull wars: Kennewick man, archaeology, and the battle for Native American identity. New York: Basic Books. Thompson, D. E., and J. V. Murra. 1966. The Inca bridges in the Huánuco region. American Antiquity 31:632–639. Thorne, A. G., and M. H. Wolpoff. 1992. The multiregional evolution of humans. Scientific American 266(4):76–83. Tobias, P. 1971. The brain in hominid evolution. New York: Columbia University Press. Todd, I. A. 1976. Çatal Hüyük in perspective. Menlo Park, CA: Cummings. Toffler, A. 1970. Future shock. London: Pan Books. Topic, T. L. 1982. The Early Intermediate period and its legacy. In Chan Chan, ed. M. E. Moseley and K. C. Day. Albuquerque: University of New Mexico Press. Topping, A. 1978. The first emperor’s army, China’s incredible find. National Geographic 153(4):440–459. Toth, N. 1987. The first technology. Scientific American 256(2):112–121. Townsend, R. F., ed. 1992. The ancient Americas: Art from sacred landscapes. Chicago: Art Institute of Chicago. Townsend, R. F., ed. 2004. Hero, hawk, and open hand: American Indian art of the ancient Midwest and South. Chicago: Art Institute of Chicago. Trask, L. 1998. The origins of speech. Cambridge Archaeological Journal 8:69–94. Trigger, B. G. 1978. Early Iroquoian contacts with Europeans. In Handbook of North American Indians. Vol. 15, Northeast, ed. W. C. Sturtevant and B. G. Trigger. Washington, DC: Smithsonian Institution Press. Trigger, B. G. 1980a. Archaeology and the image of the American Indian. American Antiquity 45:662–675. Trigger, B. G. 1980b. Gordon Childe: Revolutions in archaeology. New York: Columbia University Press.
Trigger, B. G. 1998. Sociocultural evolution: Calculation and contingency. Malden, MA: Blackwell. Trigger, B. G. 2003. Understanding early civilizations: A comparative study. Cambridge: Cambridge University Press. Trigger, B. G. 2006. A history of archaeological thought. Cambridge: Cambridge University Press. Trigger, B. G., B. J. Kemp, D. O. O’Connor, and A. B. Lloyd. 1985. Ancient Egypt: A social history. Cambridge: Cambridge University Press. Tringham, R. 1971. Hunters, fishers, and farmers of Eastern Europe 6000–3000 B.C. London: Hutchinson University Library. Trinkaus, E., ed. 1990. The emergence of modern humans. Cambridge: Cambridge University Press. Trinkaus, E. 2005. Early modern humans. Annual Review of Anthropology 34:207–230. Trinkhaus, E. 2007. European early modern humans and the fate of the Neandertals. PNAS 104:7367–7372. Trinkaus, E., and W. W. Howells. 1979. The Neanderthals. Scientific American 241(6):94–105. Trubitt, M. B. 2003. Mississippian period warfare and palisade construction at Cahokia. In Theory, method, and practice in modern archaeology, ed. R. J. Jeske and D. K. Charles. Westport, CT: Praeger. Tuck, J. A. 1978a. Northern Iroquoian prehistory. In Handbook of North American Indians. Vol. 15, Northeast, ed. W. C. Sturtevant and B. G. Trigger. Washington, DC: Smithsonian Institution Press. Tuck, J. A. 1978b. Regional cultural development, 3000 to 300 B.C. In Handbook of North American Indians. Vol. 15, Northeast, ed. W. C. Sturtevant and B. G. Trigger. Washington, DC: Smithsonian Institution Press. Turner, B. L., II, and P. D. Harrison, eds. 1983. Pulltrouser Swamp: Ancient Maya habitat, agriculture, and settlement in northern Belize. Austin: University of Texas Press. Tyler, N. 1999. Historic preservation: An introduction to its history, principles, and practice. New York: Norton. Tylor, E. B. 1960. Anthropology. Ann Arbor: University of Michigan Press. Originally published 1881.
_ References
R-21
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
Ubelaker, D. H. 1978. Human skeletal remains. Washington, DC: Taraxacum Press. Ucko, P. J., and A. Rosenfeld. 1967. Paleolithic cave art. London: Weidenfeld & Nicholson. Underhill, A. 1997. Current issues in Chinese Neolithic archaeology. Journal of World Prehistory 11:103–160. Underhill, A. P. 2002. Craft production and social change in northern China. New York: Kluwer/Plenum. Underhill, A. P., G. M. Feinman, L. M. Nicholas, G. Bennett, H. Fang, F. Luan, H. Yu, and F. Cai. 2002. Regional survey and the development of complex societies in southeastern Shandong, China. Antiquity 76:745–755. Underhill, A., G. Feinman, L. Nicholas, H. Fang, F. Luan, H. Yu, and F. Cai. 2008. Changes in regional settlement patterns and the development of complex societies in southeastern Shandong, China. Journal of Anthropological Archaeology 27:1–29. Ungar, P. S., ed. 2006. Evolution of the human diet: The known, the unknown, and the unknowable. Oxford: Oxford University Press. Urton, G. 2003. Signs of the Inka Khipu: Binary coding in the Andean knottedstring records. Austin: University of Texas Press. U.S. Congress, Office of Technology Assessment. 1986. Technologies for prehistoric and historic preservation. OTA-E-319. Washington, DC: U.S. Government Printing Office. Vaillant, G. C. 1966. Aztecs of Mexico. Harmondsworth, England: Pelican. Vermeule, E. 1972. Greece in the Bronze Age. Chicago: University of Chicago Press. Vigne, J. D., J. Guilaine, K. Debue, L. Haye, and P. Gérard. 2004. Early taming of the cat in Cyprus. Science 304:259–260. Villa, P. 1982. Conjoinable pieces and site formation processes. American Antiquity 47:276–290. Vitelli, K. D. 1996. Archaeological ethics. Walnut Creek, CA: AltaMira. Vitelli, K. D., and C. ColwellChanthphonh. 2006. Archaeological ethics, 2d ed. Blue Ridge Summit, PA: AltaMira. Wainwright, G. 1989. The henge monuments. London: Thames & Hudson.
R-22
References
Page R-22
Walker, A. 1981. Diet and teeth: Dietary hypotheses and human evolution. Philosophical Transactions of the Royal Society of London B292:57–64. Wallace, A. R. 1869. Malay archipelago. New York: Harper & Brothers. Warren, P. 1975. The Aegean civilizations. Oxford: Elsevier Phaidon. Warrick, G. A. 1983. Reconstructing Iroquoian village organization. National Museum of Man Mercury Series, Paper No. 124. Ottawa: Archaeological Survey of Canada. Warrick, G. A. 1988. Estimating Ontario Iroquoian village duration. Man in the Northeast 36:21–60. Watson, W. 1960. Archaeology in China. London: Max Parrish. Waters, M. R., and T. W. Stafford Jr. 2007. Redefining the age of Clovis: Implications for the peopling of the Americas. Science 315:1122–1126. Weaver, M. P. 1981. The Aztecs, Maya, and their predecessors: Archaeology of Mesoamerica. New York: Academic Press. Webb, C. H. 1982. The Poverty Point culture, 2d ed. Baton Rouge: Louisiana State University School of Geoscience. Webster, D. 2002. The fall of the ancient Maya. London: Thames & Hudson. Weiss, E., M. E. Kislev, and A. Hartmann. 2006. Autonomous cultivation before domestication. Science 312:1608–1610. Weiss, H. 2005. Collapse: How sudden climate change destroyed civilization and shaped history. London: Routlege. Wenke, R. J. 1999. Patterns in prehistory: Humankind’s first three million years, 4th ed. New York: Oxford University Press. Wenke, R. J. 2009. The ancient Egyptian state: The origins of Egyptian culture (c. 20,000–1900 B.C.). Cambridge: Cambridge University Press. Wheeler, R. E. M. 1943. Maiden Castle, Dorset. London: Society of Antiquaries 12. Wheeler, R. E. M. 1968. The Indus civilization. Cambridge: Cambridge University Press. Wheeler, T. S., and R. Maddin. 1976. The techniques of the early Thai metalsmith. Expedition 18(4):38–47. White, J. C. 1982. Discovery of a lost Bronze Age: Ban Chiang. Philadelphia: University of Pennsylvania Press. White, L. A. 1959. The evolution of culture. New York: McGraw-Hill.
White, R. 1986. Dark caves, bright visions: Life in ice age Europe. New York: American Museum of Natural History. White, T., G. Suwa, and B. Asfaw. 1994. Australopithecus ramidus, a new species of early hominid from Aramis, Ethiopia. Nature 371:306–312. Whittle, A. 2003. The archaeology of people: Dimensions of Neolithic life. London: Routledge. Wilcox, D. R., T. R. McGuire, and C. Sternberg. 1981. Snaketown revisited. Arizona State Museum Archaeological Series 155. Tucson: University of Arizona. Willey, G. R. 1953. Prehistoric settlement in the Virú Valley, Peru. Washington, DC: Smithsonian Institution Press. Willey, G. R. 1966. An introduction to American archaeology. Vol. 1, North and Middle America. Englewood Cliffs, NJ: Prentice-Hall. Willey, G. R. 1971. An introduction to American archaeology. Vol. 2, South America. Englewood Cliffs, NJ: Prentice-Hall. Willey, G. R. 1974. The Classic Maya hiatus: A rehearsal for the collapse? In Mesoamerican archaeology: New approaches, ed. N. Hammond. London: Duckworth. Wills, W. H. 1988. Early agriculture and sedentism in the American Southwest: Evidence and interpretations. Journal of World Prehistory 2:445–488. Wilmsen, E. N. 1974. Lindenmeier: A Pleistocene hunting society. New York: Harper & Row. Wilmsen, E. N. 1978. Lindenmeier, 1934–74. Washington, DC: Smithsonian Institution Press. Wilson, A. C., and R. L. Cann. 1992. The recent African genesis of humans. Scientific American 266(4):68–73. Wilson, D. J. 1981. Of maize and men: A critique of the maritime hypothesis of state origins on the coast of Peru. American Anthropologist 83:93–120. Wilson, S. M., ed. 1997. The indigenous people of the Caribbean. Gainesville: University Press of Florida. Wing, E. S. 1980. Faunal remains. In Guitarrero Cave: Early man in the Andes, ed. T. S. Lynch. New York: Academic Press. Wiseman, J. R., and F. El-Baz, eds. 2007. Remote sensing in archaeology. New York City: Springer.
pri31057_ref.ppR1-R24.qxd
10/2/09
8:31 AM
Wittfogel, K. 1957. Oriental despotism. New Haven, CT: Yale University Press. Wolf, E. R. 1982. Europe and the people without history. Berkeley: University of California Press. Wolpoff, M. 1999. Paleoanthropology, 2d ed. Boston: McGraw-Hill. Wood, B. Palaeoanthropology: Hominid revelations from Chad. Nature 418:133–135. Wood, B., and B. G. Richmond. 2000. Human evolution: Taxonomy and paleobiology. Journal of Anatomy 196:19–60. Wood, J. W., D. Lai, P. L. Johnson, K. L. Campbell, and I. M. Masler. 1985. Lactation and birth spacing in highland New Guinea. Journal of Biosocial Science, Supplement 9:159–173. Woodman, P. C. 1981. A Mesolithic camp in Ireland. Scientific American 245(2):120–132. Woolley, C. L. 1954. Excavations at Ur. London: Benn. Wright, G. A. 1969. Obsidian analyses and prehistoric Near Eastern trade: 7500–3500 B.C. Ann Arbor: University of Michigan. Anthropological Papers of the Museum of Anthropology, No. 37. Wright, H. E. 1971. Late Quaternary vegetational history of North America. In The Late Cenozoic ice ages, ed. K. K. Turekian. New Haven, CT: Yale University Press. Wright, H. T. 1986. The evolution of civilizations. In American archaeology, past and future, ed. D. J. Meltzer, D. D. Fowler, and J. A. Sabloff. Washington, DC: Smithsonian Institution Press. Wright, H. T. 2005. The polycentricity of the archaic civilizations. In A catalyst for ideas: Anthropological archaeology and the legacy of Douglas W. Schwartz,
Page R-23
ed. V. L. Scarborough. Santa Fe, NM: School of American Research Press. Wright, H. T., and G. A. Johnson. 1975. Population, exchange and early state formation in southwestern Iran. American Anthropologist 77:267–289. Wright, H. T., and E. S. A. Rupley. 2001. Calibrated radiocarbon age determinations of Uruk-related assemblages. In Uruk Mesopotamia and its neighbors: Cross-cultural interactions in the era of state formation, ed. M. S. Rothman. Santa Fe, NM: School of American Research Press. Wright, R. P. 2009. The ancient Indus: Urbanism, economy, and society. Cambridge: Cambridge University Press. Wymer, J. 1968. Lower Paleolithic archaeology in Britain. London: John Baker. Xi, Z., F. Zhang, B. Xu, J. Tan, S. Li, C. Li, H. Zhou, H. Zhu, J. Zhang, Q. Duan, and L. Jin. 2008. Mitochondrial DNA evidence for a diversified origin of workers building mausoleum for first emperor of China. PLoS One 10(3):1–7. Yates, R. D. S. 1994. The birth of imperial China. In Cradles of civilization: China, ed. R. E. Murowchick. Norman: University of Oklahoma Press. Yen, D. E. 1977. Hoabinhian horticulture: The evidence and the questions from northwest Thailand. In Sunda and Sahul: Prehistoric studies in Southeast Asia, Melanesia, and Australia, ed. J. Allen, J. Golson, and R. Jones. London: Academic Press. Yen, D. E. 1982. Ban Chiang pottery and rice. Expedition 24(4):51–64. Yerkes, R. W. 1988. The Woodland and Mississippian traditions in the prehistory of midwestern North America. Journal of World Prehistory 2:307–358.
Yoffee, N., and G. L. Cowgill, eds. 1988. The collapse of ancient states and civilizations. Tucson: University of Arizona Press. Zarrillo, S., D. M. Pearsall, J. S. Raymond, M. A. Tisdale, and D. J. Quon. 2008. Directly dated starch residues document Early Formative maize (Zea mays L.) in tropical Ecuador. Proceedings of the National Academy of Sciences 103(33):5006–5011. Zeder, M. A. 2006. Central questions in the domestication of plants and animals. Evolutionary Anthropology 15:105–117. Zeder, M. A., D. Bradley, E. Emshwiller, and B. D. Smith, eds. 2006. Documenting domestication: New genetic and archaeological paradigms. Berkeley: University of California Press. Zimmer, C. 2001. Evolution: The triumph of an idea. New York: HarperCollins. Zimmerman, L. J., K. D. Vitelli, and J. Hollowell-Zimmer, eds. 2003. Ethical issues in archaeology. Walnut Creek, CA: AltaMira. Zohary, D., and M. Hopf. 2000. Domestication of plants in the Old World: The origin and spread of cultivated plants in West Asia, Europe, and the Nile Valley, 3d ed. Oxford: Oxford University Press. zur Nedden, D., K. Wicke, R. Knapp, H. Seidler, H. Wilfing, G. Weber, K. Spindler, W. A. Murphy, G. Hauser, and W. Platzer. 1994. New findings on the Tyrolean “Ice Man”: Archaeological and CT-body analyses suggest personal disaster before death. Journal of Archaeological Science 21:809–818. Zvelebil, M., and P. M. Dolukhanov. 1991. The transition to farming in eastern and northern Europe. Journal of World Prehistory 5:233–278.
_ References
R-23
pri31057_ref.ppR1-R24.qxd
_ _
10/2/09
8:31 AM
Page R-24
pri31057_cre.ppC1-C6.qxd
10/2/09
8:12 AM
Page C-1
<
Photo Credits
Chapter 1 1.1, © Boxgrove Project; 1.2, Courtesy NASA, ESA and J. Hester (ASU); 1.6, Courtesy James B. Stoltman; 1.7, T.D. Price; 1.8, Courtesy Michael Kienitz; 1.9, Courtesy Lawrence Conyers; 1.10, T.D. Price; 1.13, © John W. Rick; 1.14, T.D. Price; 1.15, © Michael A. Hampshire/NGS Image Collection; 1.16, Courtesy Michael Kienitz; 1.17, T.D. Price; 1.18, Courtesy of James B. Stoltman; 1.19, Courtesy Sönke Hartz; 1.20, 1.21, Courtesy Michael Kienitz; 1.22, T.D. Price; 1.23, © Statens Historiska Museum. Photo: Soren Hallgren; 1.25, 1.29, 1.33, T.D. Price; 1.34, © Warren & Genny Garst/ Tom Stack and Associates Chapter 2 2.1, © John Reader/Photo Researchers, Inc.; 2.2, © Kevin O’Farrell Concepts; 2.3, Courtesy of Hominidae.com; 2.9, Micrographs courtesy of Dr. Frederick E. Grine, SUNY, Stony Brook; 2.10, © John Reader/Photo Researchers, Inc.; 2.12, © Institute of Human Origins; 2.13, 2.14, 2.15, 2.16, © John Reader/Photo Researchers, Inc.; 2.17, Courtesy Department of Library Services, AMNH. Photo D. Finnin/C. Chesek. Neg. #4936-300; 2.19, © John Reader/ Photo Researchers, Inc.; 2.21, Courtesy Richard Potts; 2.22, © E.E. Kingsley/Photo Researchers, Inc.; 2.27, Robert F. Sisson, © National Geographic Image Collection; 2.28, © John Reader/Photo Researchers, Inc.; 2.30, © AP/Wide World Photos Chapter 3 3.1, © Royal British Columbia Museum, Canada; 3.3, © National Museums of Kenya; 3.4, © Tom McHugh/Photo Researchers, Inc.; 3.6, Photo courtesy F. Clark Howell; 3.12, From Professor Huang Weiwen. Institute of Paleontology & Paleoanthropology, Beijing; 3.13, Courtesy Department of Library Services. American Museum of Natural History. Neg #335651; 3.14, © Dean Conger/Corbis; 3.18 (left), Staatliches Museum fur Naturkunde, Stuttgart; 3.18 (right), © Laboratoire de Préhistoire du Musée de l’Homme; 3.19, Courtesy of F. Clark Howell; 3.21, © Javier Trueba/Madrid Scientific Films; 3.22, © Maurício Antón/ Madrid Scientific Films; 3.23, © Javier Trueba/Madrid Scientific Films; 3.24, The Age of Mammals, a mural by Rudolph F. Zallinger. Copyright © 1966, 1975, 1989, 1991, 2000 Peabody Museum of Natural History, Yale University, New Haven, Connecticut, USA.; 3.25, Aaris-Sorensen and Brinch Petersen, STRIAE, 24: 111–117, photo by Geert Brovad; 3.27, Courtesy J. Desmond Clark; 3.30, © Dr. Harmut Thieme, Niedersächsisches Landesamt für Denkmalpflege, Hannover, Germany. Photo: Peter Pfarr Chapter 4 4.1, © Erich Lessing/ Art Resource, NY; 4.3, © Kenneth Garett/NGS Image Collection; 4.7, © John Reader/Photo Researchers, Inc.; 4.8, T.D. Price; 4.15, © John Reader/Photo Researchers, Inc.; 4.16, Reconstruction of Homo sapiens neanderthalensis by Jay H. Matternes, © Copyright 1981.; 4.23, T.D. Price; 4.27, © The National Museum of Denmark; 4.29, Courtesy Peter Vemming Hansen; 4.31, © Musée de l’Homme. Photo by D. Destable; 4.32, Courtesy of Archeologicky Ustav, Brno, Czech Republic; 4.35, Collection of the Institute fur Ur-and Fruhgeschichte, University of Tubingen, Germany; 4.38, Photo by Jean Vertut, Isseyles-Molineaux, France; 4.44, © Charles & Josette Lenars/Corbis; 4.45, Courtesy J.-M. Chauvet, French Ministry of Culture and Communication, Regional Direction for Cultural Affairs - Rhône-Alpes region Regional Department of Archaeology; 4.47, Courtesy Department of Library Services. American Museum of Natural History. Neg # 39686. Photo by Kirschner; 4.48, © Professor Leroi-Gourhan. From The Old Stone Age, p. 236. Weidenfeld & Nicholson Ltd.; 4.53, © Réunion des Musées Nationaux/Art Resource, NY; 4.54, © Medford Taylor/NGS Image Collection; 4.55, © Anthropology Photographic Archive, University of Auckland; 4.62, © Dr. Mike Waters, Texas A&M University; 4.63, Courtesy of the Royal Ontario Museum, Toronto, Canada; 4.66, From The National Anthropological Archives, Smithsonian Institution; 4.68, © Keren Su/Corbis; 4.69, 4.71, © Tom Dillehay; 4.72, Courtesy Vance Holliday; 4.73, From The National Anthropological Archives, Smithsonian Institution; 4.76, © National Geographic Soci-
ety Image Collection Chapter 5 5.1, © Lennart Larsen/The National Museum of Denmark; 5.3, Courtesy Tom Pleger; 5.9, T.D. Price; 5.12, © The National Museum of Denmark; 5.13, © Lennart Larsen/The National Museum of Denmark; 5.17, T.D. Price; 5.19, Courtesy John Parkington, Department of Archaeology, University of Capetown; 5.22, 5.23, 5.24, Courtesy Sannai-Maruyama Site Preservation Office, Cultural Properties Protection Division, Aomori Prefectural Board of Education; 5.25, © Board of Trustees, Southern Illinois University; 5.27, Courtesy Richard Jeffries; 5.28, © Southern Illinois University; 5.30, © Board of Trustees, Southern Illinois University; 5.32, © Michael A. Hampshire/NGS Image Collection; 5.34, © Lee/Anthro-Photo Chapter 6 6.1, © Biblioteca Medicea-Laurenziana, Florence, Italy/The Bridgeman Art Library; 6.2, © Erich Lessing/Art Resource, NY; 6.4, T.D. Price; 6.6, Reprinted by permission from the Florentine Codex: General History of the Things of New Spain by Fray Bernardino de Sahagun, Book IV, Which Telleth of the Book of Days Which the Mexicans Handed Down, translated by Arthur J.O. Anderson and Charles E. Dibble. © 1981 by the School of American Research, Santa Fe.; 6.8, Photo by Peter Dorell and Stuart Laidlaw, Courtesy University of London, Institute of Archaeology; 6.9, Courtesy of James L. Phillips, The Field Museum; 6.16, Courtesy Charles M. Niquette, Cultural Resource Analysts, Inc.; 6.17, © Gil Stein; 6.18, Courtesy Institute of Archaeology, University College, London; 6.20, © Albatross Aerial Photography; 6.21, © Ann Hatfield; 6.23, © Barry Kass/Images of Archaeology; 6.29, Courtesy Ian Hodder, Catalhöyük Project, Stanford University; 6.31, Courtesy James Mellaart; 6.34, © C. Jarrige/CNRSGuimet; 6.37, Courtesy Susan Kepecs; 6.39, Courtesy Linda Nicholas; 6.46, 6.47, 6.48, © Charles Higham; 6.51, Courtesy Kent V. Flannery; 6.52, Courtesy Linda Nicholas; 6.53, Courtesy Kent V. Flannery; 6.54, © David Cavagnaro/Peter Arnold, Inc.; 6.57, 6.58, © Robert S. Peabody Museum of Archaeology. Phillips Academy, Andover, MA. All rights reserved.; 6.62, 6.63, 6.64, 6.65, © Thomas F. Lynch; 6.66, © Chip Clark; 6.68, T.D. Price Chapter 7 7.1, Courtesy Cahokia Mounds State Historic Site, Collinsville, Illinois; 7.4, © Richard A. Cooke/Corbis; 7.6, Courtesy Poverty Point State Historic Site, Louisiana Office of State Parks. Photo by Steven Carricut; 7.7, © The Field Museum, Neg. #94855; 7.12, © The Field Musuem, A113968. Photo by John Weinstein; 7.13, © The Field Museum, A110016C. Photo by Ron Testa; 7.14, Panther Effigy Pipe, USA, Indiana, Posey County, Mann Site. AllisonCopena, Middle Woodlands period, AD 1-400. Black steatite. 2 3/8 x 1 9/16 x 6 5/16, L49.5. Anonymous loan. © The Brooklyn Museum of Art; 7.15, © The Field Museum, A113969. Photo by John Weinstein; 7.16, © The Field Museum, Neg. #A110028c; 7.17, Photo © Sam Noble Oklahoma Museum of Natural History, University of Oklahoma. Used with permission of the Caddo Nation and the Witchita & Afflitated Tribes; 7.19, Courtesy Cahokia Mounds State Historic Site, Collinsville, Illinois. Painting by William R. Iseminger; 7.20, Courtesy Dr. Melvin Fowler, University of Wisconsin, Milwaukee; 7.21, Courtesy Cahokia Mounds State Historic Site, Collinsville, Illinois; 7.22, Courtesy Missouri Historical Society, St. Louis; 7.24, Library, Academy of Natural Sciences, Philadelphia; 7.31, Photo by William D. Finlayson/This Land Archaeology, Inc.; 7.32, © The Field Museum, Neg. #CSA78062; 7.34, Arizona State Museum, The University of Arizona. Photo by Helga Teiwes; 7.35, Arizona State Museum, The University of Arizona. Photo by Emil Haury; 7.36, © The Field Museum, cat. #83350. Photograph by Linda Nicholas; 7.37, Arizona State Museum, The University of Arizona. Photo by Helga Teiwes; 7.38, Courtesy the Arizona State Museum, the University of Arizona, photographer, Jannelle Weakly. #7301; 7.42, Courtesy Linda Nicholas; 7.43, © The Field Museum, Neg. #CSA77994; 7.44, © The Field Museum, Neg. #A88593; 7.45, Courtesy Paul Logsdon; 7.47, © Ruth &
Credits
C-1
_
pri31057_cre.ppC1-C6.qxd
10/15/09
11:14 AM
Page C-2
Louis Kirk; 7.48, © Ruth Kirk; 7.52, Courtesy Allan May; 7.53, © Birmingham Public Library, Cat. #245; 7.54, © The Field Museum, Neg. #A97603 Chapter 8 8.1, Photo: © David Hiser/NGS Image Collection, Overlay: Ned M. Seidler, © National Geographic Society; 8.3, 8.7, Courtesy Linda Nicholas; 8.9, Richard H. Stewart, © National Geographic Society Image Collection; 8.11, © Museo Nacional de Antropologia, Mexico City, Mexico/Bridgeman Art Library/Giraudon; 8.14, Courtesy Department of Library Services, American Museum of Natural History. Neg #1298; 8.16, 8.19, Courtesy Dr. Ray T. Matheny/ Project El Mirador; 8.25, 8.27, Courtesy Linda Nicholas; 8.28, Courtesy of Stephen Kowalewski/University of Georgia; 8.32, 8.33, Courtesy Linda Nicholas; 8.36, Courtesy Jennifer Blitz; 8.39, © Arvind Garg/Corbis; 8.40, © Charles & Josette Lenars/Corbis; 8.41, © University of Pennsylvania Museum, Tikal Project, Neg. #65-4-1094; 8.44, From Science, “ Pre-Historic Raised Field Agriculture in the Maya Lowlands,” by B.L Turner, Volume 213, pp. 399–405, July 1981. © 1981 by the AAAS; 8.48, © Werner Foreman/Art Resource, NY; 8.49, 8.55, Courtesy Linda Nicholas; 8.57, Courtesy Susan Kepecs; 8.58, © Bates Littlehales/NGS Image Collection; 8.60, 8.61, Courtesy Linda Nicholas; 8.62, © The Field Museum, Neg. #CSA45113; 8.64, © Kenneth Garrett; 8.68, © The Bodleian Library, Oxford University. MS. Arch Seld. A.1, fol. 47r; 8.71, Codex Florentino: Illustrations for Sahagun’s Historia General de las Cosas de Nueva España, edited by Francisco del Paso Troncoso, Vol. 5, Madrid, 1905.; 8.73, © Boltin Picture Library/ Bridgeman Art Library Chapter 9 9.1, © Nathan Benn; 9.3, © Jonathan Haas/Getty Images; 9.4, R.A. Feldman, Ph.D dissertation, Harvard University; 9.5, 9.8, Courtesy Jeffrey Quilter; 9.10, © Johan Reinhard; 9.12, © Wilfredo Loayza; 9.14, © John W. Rick; 9.16, Courtesy Dumbarton Oaks Research Library and Collections, Washington, DC; 9.17, Courtesy Department of Library Services. American Museum of Natural History. Neg. #334913; 9.21, Courtesy Department of Library Services. American Museum of Natural History. Neg #122461; 9.23, © William Albert Allard/NGS Image Collection; 9.24, © Bill Ballenberg/NGS Image Collection; 9.25, Photo by Guillermo Cock; 9.27, © Nathan Benn; 9.28, © Martha Cooper/NGS Image Collection; 9.31, © 2006 Peabody Museum, Harvard University. 50-1-20/22982 T268; 9.32, Courtesy Richard F. Townsend; 9.37, Courtesy Michael E. Moseley; 9.38, Courtesy Department of Library Services. American Museum Natural History. Neg #334878; 9.41, Courtesy Jennifer Blitz; 9.42, © Robert Frerck/Woodfin Camp and Associates; 9.44, Courtesy John Hyslop, American Museum of Natural History; 9.48, Courtesy Donald E. Thompson Chapter 10 10.4, © John Sanders of the The Oriental Institute, Chicago; 10.9, © Hirmer Verlag; 10.10, © Gil Stein/ Oriental Institute, University of Chicago; 10.11, Courtesy Denise Schmandt-Besserat. With permission of Deutsches Archaologisches Institut; 10.12a, © University of Pennsylvania Museum (Image #141589); 10.12b, © University of Pennsylvania Museum (Image #142672); 10.12c, © University of Pennsylvania Museum (Image #141589); 10.13, 10.14, Courtesy Denise Schmandt-Besserat. With permission of Deutsches Archaologisches Institut; 10.16, © Dilip Mehta/Woodfin Camp and Associates; 10.19, Courtesy George Dales and the Harappa Project, University of California at Berkeley; 10.20, Photo by J.M. Kenoyer, courtesy of the Department of Archaeology and Museums,
_ _
C-2
Credits
Government of Pakistan; 10.21, Courtesy George Dales and the Harappa Project, University of California at Berkeley; 10.24, © Ashmolean Museum, Oxford; 10.25, © Erich Lessing/Art Resource, NY; 10.29, © Sandro Vannini/Corbis; 10.30, © Photri, Inc.; 10.32, Acct. no. 11.1738, King Menkaure (Mycerinus) and Queen. Harvard University/ Museum of Fine Arts Expedition. Photograph © Museum of Fine Arts, Boston; 10.40, Courtesy of the Institute History and Philology, Academy Sinica, Taiwan; 10.41, Courtesy the Metropolitan Museum of Art, Gift of Mrs. John Marriott, Mrs. John Barry Ryan, Gilbert W. Kahn, Roger Wolfe Kahn, 1949. #49. 136.5; 10.42, Photo courtesy Mr. Ye Maolin. Reprinted by permission from Macmillan Publishers Ltd: Nature, Vol. 437, Issue 7061. Copyright 2005; 10.44, © Audrey R. Topping; 10.45, 10.47, Courtesy Linda Nicholas; 10.48, © Asian Art & Archaeology, Inc./Corbis; 10.50, Courtesy NASA/JPL/Caltech; 10.51, Glen Allison/Getty Images/PhotoDisc; 10.52, © Luca I. Tettoni/ Corbis; 10.54, 10.55, © Dr. Roderick James McIntosh; 10.56, © The Granger Collection, New York; 10.60, © Robert Holmes; 10.61, © Walter Edwards/NGS Image Collection Chapter 11 11.1, © Scala/Art Resource, NY; 11.3, © 1991 by Peter Bogucki. Used by permission; 11.5, © Natalie Munro; 11.6, Courtesy Thomas Jacobsen; 11.7, 11.8, Courtesy of the Museum of Archaeology in Varna, Bulgaria; 11.9, © V. Vitanov/Ancient Art & Architecture Collection; 11.10, Forschungsinstitut für Alpine Vorzeit; 11.11, © Gerha Hinterleitner/ Gamma/Eyedea; 11.14, Courtesy Musee Dauphinois. Grenoble, France; 11.17, 11.18, A. Bocquet/Centre de Documentation de la Prehistoire Alpine; 11.20, T.D. Price; 11.21, From Ancient Europe from the Beginnings of Agriculture to Classical Antiquity, by Stuart Piggott. Reprinted with permission from Edinburgh University Press; 11.23, © British Crown Copyright/ M.O.D. Reproduced with the permission of the Controller of Her Britannic Majesty’s Stationary Office; 11.29, © Vincenzo Lombardo/Taxi/Getty Images; 11.31, Courtesy John Bennet, Oxford University, Department of Classics; 11.33, © Kevin Fleming/Corbis; 11.34, Courtesy Walter M. Shandruk; 11.35, Photo by J. McCredie, National Archaeological Museum, Athens; 11.36, © Gianni Dagli Orti/Corbis; 11.39, Four bronze age barrows along the coast of Denmark. From Danish Prehistoric Monuments, by P.V. Glob, 1971, page 27, figure 4. Reprinted with permission of Glydendal, Copenhagen; 11.40, © Lennart Larsen/The National Museum of Denmark; 11.43, From J. Jensen, Danmarks oldtid bd. 2, Bronzealderen 2000-500 f.Kr. 2002 Gyldendal. København. Used by permission of the author; 11.44, © The National Museum of Denmark; 11.47, Musee Archeologique, Chatillon Sur Seine, France/Giraudon/ Bridgeman Art Library; 11.46, Cliche Musee Archeologique de Chatillon sur Seine; 11.48, © Forhistorisk Museum, Moesgard; 11.49, © Ira Block; 11.50, © Dae Sasitorn/www.lastrefuge.co.uk; 11.52, Courtesy the Society of Antiquaries of London; 11.54, © UK Alan King/Alamy Chapter 12 12.1, K.L Feder; 12.2, Digital Vision/Getty Images; 12.3, Jeremy Woodhouse/PhotoDisc/Getty Images; 12.4, © Reuters/Corbis; 12.5, Photo by Time Inc./Time Life Pictures; 12.6, © Jonathan Blair/ Corbis; 12.7, Poster courtesy of the State Historic Preservation Office, Arizona State Parks, Phoenix, Arizona; 12.9, © Diego Meozzi/Stone Pages
pri31057_cre.ppC1-C6.qxd
10/15/09
12:35 PM
Page C-3
<
Text Credits
Chapter 1 1.12, Courtesy of Goran Burenhult; 1.26, Courtesy of Sonke Hartz; 1.28, Adapted from A. Sherratt (ed.), Cambridge Encyclopedia of Archaeology. Copyright © 1980 Cambridge University Press; 1.31, Courtesy of Ullrich Rossing; 1.32, With permission from Annual Review of Ecology and Systematics, Vol. 3. © 1972 by Annual Reviews: www.annualreviews.org Chapter 2 2.4, Kottak, C., Anthropology: The Exploration of Human Diversity, 9/e, fig. 6.3, p. 146. Copyright © The McGraw-Hill Companies, Inc. Used with permission from The McGraw-Hill Companies; 2.6, From P. Gagneux et al., “Mitochondrial Sequences Show Diverse Evolutionary Histories of African Hominids,” Proceedings of the National Academy of Sciences, Vol. 96, April 27, 1999, pp. 5077–5082. Copyright © 1999 National Academy of Sciences, U.S.A. Used with permission; 2.7, From Pat Shipman, “Hunting the First Hominid,” American Scientist, Vol. 90, No. 1, Jan–Feb 2002, p. 27. Illustration credit: Barbara Aulicino/American Scientist. Copyright © 2002 American Scientist. Reprinted by permission of American Scientist, magazine of Sigma Xi, The Scientific Research Society; 2.11, From Johansen & Edey, Lucy: The Beginnings of Humankind, Simon & Schuster; 2.18, From footprints on PBS website; 2.25, Mary Leakey, Olduvai Gorge, Vol. 3. Copyright © Cambridge University Press; 2.30, Nicholas Toth, “The First Technology,” Scientific American, April 1987, art by Edward L. Hanson. Art copyright © Edward L. Hanson. Used by permission of the artist Chapter 3 3.7, From Andre LeroiGourhan, 1957, Prehistoric Man. New York: Philosophical Library. Reprinted with permission; 3.24, The Age of Mammals, a mural by Rudolph F. Zallinger. Copyright © 1966, 1975, 1989, 1991, 2000 Peabody Museum of Natural History, Yale University, New Haven, CT; 3.26, Courtesy of Richard Klein; 3.28, Adapted from J. Desmond Clark, Kalambo Falls Prehistoric Site, Vol. 1, Cambridge University Press, 1969; 3.29, From Francois Bordes, The Old Stone Age. Copyright © 1968 The McGraw-Hill Companies, Inc. Used with permission from The McGraw-Hill Companies Chapter 4 4.6, From Roger Lewin, In the Age of Mankind, (Washington, DC: Smithsonian Institution Press), page 181. Used by permission. Copyright © 1989; 4.10, Courtesy of Richard Klein; 4.12, From Mosaic, Vol. 10, No. 2, March/April 1979; 4.13, Feder, K. and M. Park, Human Antiquity: An Introduction to Physical Anthropology and Archaeology, 4/e, f4.1, p. 71. Copyright © The McGraw-Hill Companies, Inc. Used with permission from The McGraw-Hill Companies; 4.18, From Francois Bordes, The Old Stone Age. Copyright © 1968 The McGraw-Hill Companies, Inc. Used with permission from The McGraw-Hill Companies; 4.19, Kottak, C., Anthropology: The Exploration of Human Diversity, 9/e, fig. 7.4, p. 176. Copyright © The McGraw-Hill Companies, Inc. Used with permission from The McGraw-Hill Companies; 4.20, From Turnbaugh et al., Understanding Physical Anthropology & Archaeology, 7/e, p. 439. Copyright © 1999. Reprinted with permission of Wadsworth, a division of Cengage; www.cengage.com.; 4.21, © John Sibbick/NHMPL. Used with permission; 4.24, “Out on a Limb” from P. Kahn & A. Gibbons, “Anthropology: DNA From an Extinct Human,” Science, Vol. 277, July 10, 1997 pp. 176–178. Copyright © 1997 AAAS. Reprinted with permission of Svante Pääbo, MPI-EVA; 4.25, From Oxford Illustrated Prehistory of Europe by Barry Cunliffe. Copyright © 1994. Reprinted by permission of Paul Mellars; 4.33, From B. Kilma, 1963, Dolni Vestonice, Prague: Academy of Science of the Czech Republic. Reproduced with permission; 4.34, Feder, K. and M. Park, Human Antiquity: An Introduction to Physical Anthropology and Archaeology, 4/e, f13.5, p. 395. Copyright © The McGraw-Hill Companies, Inc. Used with permission from The McGraw-Hill Companies; 4.36, H. Muller-Karpe, Handbuch Vor Vorgeschichte, 1966. C.H. Beck; 4.37, From B. Kilma, 1963, Dolni Vestonice, Prague: Academy of Science of the Czech Republic. Reproduced with permission; 4.40, From A. Leroi-Gourhan, 1982. The Dawn of European Art: An Introduction to Paleolithic Cave Painting. New
York: Cambridge University Press. Reprinted with permission of Cambridge University Press; 4.41, Copyright © Zdenek Burian. Used by permission of Jiri Hochman; 4.42, Secrets of the Ice Age by Evan Hadingham, 1979, Walker & Co. Reprinted by permission of Walker & Co.; 4.47 Drawing by Ann Hatfield, from Plato Prehistorian by Mary Settegast. Illustrations copyright Lindisfarne Books. Reprinted with permission; 4.49, From A. Leroi-Gourhan, The Old Stone Age, Cambridge University Press; 4.50, After Leroi-Gourhan and Brezillion, 1966. Reprinted with permission of the publisher, Editions du Centre National de la Recherche Scientifique, Paris; 4.51, © 2006 President and Fellows of Harvard College, Peabody Museum, Alexander Marshack; 4.52, © 2006 President and Fellows of Harvard College, Peabody Museum, Alexander Marshack; 4.58, Bruce D. Smith, the Emergence of Agriculture, WH Freeman, 1995, p. 39. Used with permission; 4.61, © Estate of Bunji Tagawa. Used with permission; 4.64, Anderson & Faugt, Archaeology of Eastern North America, Vol. 26, pp. 163–187. Copyright © 1998. Used by permission of Eastern States Archaeological Federation; 4.67, Reprinted from Environments & Extinctions, Mead, Jim I. and Meltzer, David J., 1985, with permission from the Center for the Study of the First Americans; 4.74, C. Kottak, Anthropology: The Exploration of Human Diversity, 9/e, fig 9.6, Copyright © The McGraw-Hill Companies, Inc. Used with permission from The McGraw-Hill Companies 4.75, D.H. Thomas, Archaeology, f2.1, p. 64. Copyright © 1979. Reprinted with permission of Wadsworth, a division of Cengage: www.cengage.com; Chapter 5 5.7, Legge & RowleyConway, Star Carr Revisited, 1990. Copyright © Anthony Legge. Used with permission; 5.8, Courtesy of Lars Larsson, Institute of Archaeology, Lund University, Sweden; 5.12, Illustration by Eric Claudell. Courtesy of Erik Brinch Petersen, Institute of Archaeology, University of Copenhagen; 5.18, Courtesy of Richard Klein; 5.19, Reprinted with permission from John Parkington, University of Capetown, South Africa; 5.20, Drawing adapted from the original of Tatsuo Kobayashi in Image and Life: 50,000 Years of Japanese Prehistory by Gordon Miller, Museum Note No. 5, Illustration by Gordon Miller. Copyright © Gordon Miller. Used with permission; 5.21, Reprinted from The Prehistory of Japan, by C. Aikens & T. Higuchi, Copyright © 1982. With permission from Elsevier; 5.26, Figure 1-2, p. 9, from The Archaeology of Carrier Mills: 10,000 Years in the Saline Valley of Illinois, by Richard W. Jefferies, 1987. Illustrated by Thomas W. Gatlin. © Board of Trustees, Southern Illinois University. Reprinted with permission; 5.28, Figure 3-6, p. 84, from The Archaeology of Carrier Mills: 10,000 Years in the Saline Valley of Illinois, by Richard W. Jefferies, 1987. Illustrated by Thomas W. Gatlin. © Board of Trustees, Southern Illinois University. Reprinted with permission; 5.29, Figure from E. J. Bassett, “Osteological Analysis of Carrier Mills Burials.” In The Carrier Mills Archaeology Project: Human Adaptation in the Saline River Valley, Illinois, ed. R. W. Jeffries and B. M. Butler, 2, pp. 1027–1114. Cent. Archaeolo. Invest. Res. Pap. 33. Carbondale, IL: Center for Archaeological Investigations, 1982. Used with permission of the Center for Archaeological Investigations. Copyright 1982 by the Board of Trustees, Southern Illinois University Chapter 6 6.3, Feder, K. & M. Park, Human Antiquity 4/e, f14.1, p. 437. Copyright © 2001 by The McGraw-Hill Companies, Inc. Used with permission from The McGraw-Hill Companies; 6.7, R.F. Sage, “Was Low Atmospheric CO2 During the Pleistocene a Limiting Factor for Origin of Agriculture?,” Global Change Biology, fig. 2, pp. 93–106, 1995. Copyright © 1995 Blackwell Publishing. Used with permission; 6.10, After Jean Perrot, Centre de Recherche Francais de Jerusalem; 6.11, The Rise of Civilization by Charles Redman. Copyright © 1978 by W.H. Freeman and Company. Used with permission; 6.13, Village on the Euphrates: The Excavation of Abu Hureyra edited by Andrew M.T. Moore and A. Legge. Copyright © 1999 by Oxford University Press, Inc. Used by permission of Oxford University Press; 6.14, Fig 13.14 from Credits
C-3
_
pri31057_cre.ppC1-C6.qxd
_ _
10/15/09
11:14 AM
Page C-4
Principles of Archaeology, Price. McGraw-Hill. Adapted from Figures 14.1 and 14.4 from A.M.T. Moore, G.C. Hillman, and A.J. Legge, Village on the Euphrates: From Foraging to Farming at Abu Hureyra. Oxford University Press, 2000, pp. 484, 498. Used with permission; 6.15, Reprinted with permission from the artist, Jonathan Mabry; 6.19, Courtesy Institute of Archaeology, University College London. And from Feder, K. & M. Park, Human Antiquity 4/e, f14.23, p. 475. Copyright © 2001 by The McGraw-Hill Companies, Inc. Used with permission from The McGraw-Hill Companies; 6.21, Drawing by Ann Hatfield, from Plato Prehistorian by Mary Settegast. Illustrations copyright Lindisfarne Books. Reprinted with permission; 6.22, Courtesy Institute of Archaeology, University College, London; 6.26, Clifford Jolly & White, Physical Anthropology and Archaeology 5/e. Copyright © 1995 The McGraw-Hill Companies, Inc. Used with permission from The McGraw-Hill Companies; 6.27, Legge & Rowley-Conway, Star Carr Revisited, 1990. Copyright © Anthony Legge. Used with permission; 6.28, Figure 1 from Antiquity, December 2001, p. 718. Reprinted with permission from Antiquity Publications Ltd; 6.30, Drawing by Ann Hatfield, from Plato Prehistorian by Mary Settegast. Illustrations copyright Lindisfarne Books. Reprinted with permission; 6.32, Drawing by Ann Hatfield, from Plato Prehistorian by Mary Settegast. Illustrations copyright Lindisfarne Books. Reprinted with permission; 6.33, James Mellaart, Catal-Huyuk: A Neolithic Town in Anatolia, 1966; 6.35, M. Lechevallier and G. Quivron, 1985, “Results of the Recent Excavations at the Neolithic Site Mehrgarh, Pakistan,” South Asian Archaeology, 1983. J. Schotmans and M. Taddei, eds., Instituto Universitario Orientale, Seminario di Studi Asiatici, Series Minor 23, Naples; 6.55, Feder, K. & M. Park, Human Antiquity 4/e, f14.23, p. 475. Copyright © 2001 by The McGraw-Hill Companies, Inc. Used with permission from The McGraw-Hill Companies; 6.57, Figure redrawn from G. Beadle, “The Ancestry of Corn,” Scientific American, 242:113. Art by Nelson Prentiss. © 1980. Reprinted with permission; 6.59, From “The Origins of New World Civilizations” by Richard MacNeish, Scientific American, 1964. Illustration by Eric O. Mose. Used by permission of the Estate of Eric Mose; 6.67, Reprinted by permission from Economic Botany, Vol. 35, Issue 3, p. 235, Hugh D. Wilson. Copyright © 1981 The New York Botanical Garden Bronx, New York Chapter 7 7.3, From Prehistory of North America 2/e by Jesse Jennings. Copyright © Jesse Jennings. Used by permission of the Estate of Jesse Jennings; 7.4, Adapted from Journal of Archaeological Method and Theory, Vol. 12, No. 4, December 2005, “Poverty Point as Structure, Event, Process” by Kenneth E. Sassaman (© 2005), with kind permission from Springer Science and Business Media and the author; 7.8, From Clarence Webb, The Poverty of Culture, 2/e, p. 51. Reprinted with permission from LSU Geoscience Publications; 7.18, Fig. 1.5 from Cahokia: Domination and Ideology in the Mississippian World. Copyright © 1997. Used courtesy of the Illinois Transportation Archaeological Research Program, University of Illinois at Urbana-Champaign; 7.25, Adapted from Archaeology of the Moundville Chiefdom, ed. by Vernon James Knight and Vincas P. Steponaitis, Fig. 1.1, p. 3. Smithsonian Institution Press. Used with permission. Also adapted from George R. Holley: Moundville, fig 2.6, p. 26 from Great Towns and Regional Politics, Jill Neitzel, ed. Copyright © 1999 University of New Mexico Press. Used with permission; 7.26, Redrawn from C.B. Moore, “Certain Aboriginal Remains of the Black Warrior River,” Journal of the Academy of Natural Science of Philadelphia, Vo. 13, pp. 125–244, 1905; 7.27, Steven Patricia’s rendering of Moundville. Used with permission; 7.29, Artist’s reconstruction of the main palisaded village and enclosed longhouses at the Draper site, courtesy Canadian Museum of Civilization, artist Ivan Kocsis, Mercury Series, no. 130. Used with permission; 7.31, Excavated longhouses at the Draper site, courtesy Canadian Museum of Civilization, artist Ivan Kocsis, Mercury Series, no. 130. Used with permission; 7.45, D.H. Thomas, Archaeology, 2/e, f.9.2. Copyright © 1989. Reprinted with permission of Wadsworth, a division of Cengage: www.cengage.com; 7.48, Illustration by Chris Walsh Heady from Archaeology in Washington by Ruth Kirk with Richard D. Daugherty. Copyright © 2007. Used with permission; 7.50, Illustration by Chris Walsh Heady from Archaeology in Washington by Ruth Kirk with Richard D. Daugherty. Copyright © 2007. Used with permission; 7.51, Illustration by Chris Walsh Heady from Archaeology in Washington by
C-4
Credits
Ruth Kirk with Richard D. Daugherty. Copyright © 2007. Used with permission Chapter 8 8.4, K.V. Flannery & J. Marcus, eds. The Cloud People: Divergent Evolution of the Zapotec and Mixtex Civilizations. Copyright © 1983. Courtesy of Joyce Marcus & Kent V. Flannery. Reprinted with permission; 8.5, K.V. Flannery & J. Marcus, eds. The Cloud People: Divergent Evolution of the Zapotec and Mixtex Civilizations. Copyright © 1983. Courtesy of Joyce Marcus & Kent V. Flannery. Reprinted with permission; 8.6, K.V. Flannery & J. Marcus, eds. The Cloud People: Divergent Evolution of the Zapotec and Mixtext Civilizations. Copyright © 1983. Courtesy of Joyce Marcus & Kent V. Flannery. Reprinted with permission; 8.8, K.V. Flannery & J. Marcus, eds. The Cloud People: Divergent Evolution of the Zapotec and Mixtext Civilizations. Copyright © 1983. Courtesy of Joyce Marcus & Kent V. Flannery. Reprinted with permission; 8.10, From Rebecca Gonzalez Lauck, “La Venta: An Olmec Capital” in Olmec Art of Ancient Mexico, Elizabeth Benson and Beatriz de la Fuente, eds., National Gallery of Art, Washington D.C. Copyright © Rebecca Gonzalez Lauck. Reprinted with permission; 8.14, K.V. Flannery & J. Marcus, eds. The Cloud People: Divergent Evolution of the Zapotec and Mixtext Civilizations. Copyright © 1983. Courtesy of Joyce Marcus & Kent V. Flannery. Reprinted with permission; 8.15, Drawing from In the Land of the Olmec: The Archaeology of San Lorenzo Tenochtitlan, Volume I, by Michael Coe and Richard Diehl, Copyright © 1980. By permission of the University of Texas Press; 8.17, © T.W. Rutledge; 8.20, K.V. Flannery & J. Marcus, eds. The Cloud People: Divergent Evolution of the Zapotec and Mixtext Civilizations. Copyright © 1983. Courtesy of Joyce Marcus & Kent V. Flannery. Reprinted with permission; 8.21, K.V. Flannery & J. Marcus, eds. The Cloud People: Divergent Evolution of the Zapotec and Mixtext Civilizations. Copyright © 1983. Courtesy of Joyce Marcus & Kent V. Flannery. Reprinted with permission; 8.23, K.V. Flannery & J. Marcus, eds. The Cloud People: Divergent Evolution of the Zapotec and Mixtext Civilizations. Copyright © 1983. Courtesy of Joyce Marcus & Kent V. Flannery. Reprinted with permission; 8.24, K.V. Flannery & J. Marcus, eds. The Cloud People: Divergent Evolution of the Zapotec and Mixtext Civilizations. Copyright © 1983. Courtesy of Joyce Marcus & Kent V. Flannery. Reprinted with permission; 8.25, K.V. Flannery & J. Marcus, eds. The Cloud People: Divergent Evolution of the Zapotec and Mixtext Civilizations. Copyright © 1983. Courtesy of Joyce Marcus & Kent V. Flannery. Reprinted with permission; 8.26, “A Fuego y Sangre: Early Zapotec Imperialism in Cuicatlan,” by Elizabeth Redmond, Memoirs, no. 16. Museum of Anthropology, University of Michigan, 1983. Used with permission; 8.35, Arthur G. Miller, fig. 173, p. 100 from The Mural Painting of Teotihuacan, with drawings by Felipe Davalos G. Copyright © Dumbarton Oaks. Used with permission; 8.37, Figure 23 from The Sculpture of Elajin, Veracruz, Mexico by Michael Kampen. Reprinted with permission of the University Press of Florida; 8.42, Description: Stella 29, Credit: Reprinted from John S. Henderson: The World of the Ancient Maya–Second Edition. Copyright © 1981 by Cornell University. Used by permission of the publishers, Cornell University Press and John Murray, Ltd.; 8.43, Reprinted from John S. Henderson: The World of the Ancient Maya–Second Edition. Copyright © 1981 by Cornell University. Used by permission of the publishers, Cornell University Press and John Murray Ltd.; 8.47, © Merle Greene Robertson. Used by permission; 8.50, Courtesy Andromeda Oxford Ltd., now part of the Brown Reference Group plc.; 8.51, Courtesy Andromeda Oxford Ltd., now part of the Brown Reference Group plc.; 8.52, From Everyday Life of the Aztecs by Warwick Bray, 1968. B.T. Batsford, Ltd., London; 8.55, From Tula: The Toltec Capital of Ancient Mexico by Dr. Richard A. Diehl, published by Thames and Hudson. Reprinted with permission; 8.56, Reprinted with permission from Tula of the Toltecs by Dan Healan, published by the University of Iowa Press; 8.60, Morley, Brainerd, and Sharer, The Ancient Maya, 3rd Edition. Copyright © 1946, 1947, 1956, 1983, 1994 by the Board of Trustees of the Leland Stanford Junior University. All rights reserved. Used with the permission of Stanford University Press: www.sup.org; 8.65, From Everyday Life of the Aztecs by Warwick Bray, 1968. B.T. Batsford, Ltd., London; 8.67, Frances Berdan et al., Aztec Imperial Strategies, fig. 11.1, p. 112. Copyright © 1996 Dumbarton Oaks. Used with permission; 8.68, From Everyday Life of the Aztecs by Warwick Bray, 1968. B.T. Batsford, Ltd., London; 8.70, Codex Florentino: Illustrations For Sahagen’s Historia
pri31057_cre.ppC1-C6.qxd
10/15/09
11:14 AM
Page C-5
General De Las Cosas De Nueva Espana, Francisco Del Pas Troncose, Vol. 5, Madrid, 1905; 8.71, F. Berdan, The Aztecs of Central Mexico: An Imperial Society, f. 5.3. Copyright © 1982. Reprinted with the permission of Wadsworth, a division of Cengage: www.cengage.com; 8.72, F9.6b, p. 222 from Michael Smith, The Aztecs. Copyright © 1996 Michael E. Smith. Published by Blackwell Publishing, Ltd. Used with permission Chapter 9 9.6, Chavin de Huantar, R.A. Feldman doctoral dissertation; 9.7, Copyright © Bernardino Ojeda. Used with permission; 9.9, Michael K. Mosely, Ocean Currents, The Maritime Foundations of Andean Civilization. Copyright © 1975 Michael Moseley. Reprinted with permission; 9.11, Michael K. Moseley, Ocean Currents, The Maritime Foundations of Andean Civilization. Copyright © 1975 Michael Moseley. Reprinted with permission; 9.13, Images from p. 131, Chavin and the Origins of Andean Civilization by Richard L. Burger. © 1992 Thames and Hudson, Ltd., London. Reprinted with permission; 9.15, G. R. Willey Introduction to American Archaeology: South America. © 1971 Gordon R. Willey; G. R. Willey, Introduction to American Archaeology: South America. © 1971 Gordon R. Willey; 9.18, From American Antiquity, Vol. 40, 1975. Reprinted by permission of the Society for American Archaeology; 9.19, From “Prehistoric Principles of Labor Organization in the Moche Valley, Peru,” by Michael E. Moseley and Charles Hastings, American Antiquity, 40(2): 191–196, 1975. Reprinted with permission; 9.22, Christopher Donnan, Moche Art and Iconography, Los Angeles: UCLA Latin American Center Publications, 1976. Reproduced with permission of The Regents of the University of California; 9.26, Courtesy of Christopher Donnan, drawing by Alberto Gutierrez. Reprinted with permission from Walter Leonel Alva Alva; 9.29 Courtesy Andromeda Oxford Ltd., now part of the Brown Reference Group plc.; 9.33, W. Bennett, 1947. Handbook of South American Indians, Vol. 2, J. Stewart, ed. US Government Printing Office; 9.35, Courtesy Andromeda Oxford Ltd., now part of the Brown Reference Group plc.; 9.36, Courtesy Andromeda Oxford Ltd., now part of the Brown Reference Group plc.; 9.40 From Prehistory of North America, 2/e by Jesse Jennings. Copyright © Jesse Jennings. Used by permission of the Estate of Jesse Jennings; 9.46, From Guaman Poma de Ayala, Noevo Coronica, c. 1610; 9.47, Adapted from R.W. Keating, ed., Peruvian Prehistory: An Overview of Pre-Inca and Inca Society. Copyright © 1988 Cambridge University Press; 9.49, From Craig Morris and Donald E. Thompson, Huanuco Pampa: An Inca City and its Hinterland, 1985. London: Thames and Hudson. Reproduced with permission of the authors Chapter 10 10.6, From Art of the Ancient Near East by Seton Lloyd. Copyright © 1961 Thames & Hudson, Ltd., London. Reprinted with permission; 10.7, © State Antiquities and Heritage Organization, Baghdad; 10.17, M. Wheeler, The Indus Civilization. Copyright © 1968 Cambridge University Press. Reprinted with permission; 10.18, From Michael Jansen, Civilizations Anciennes Du Pakistan, 1989. Bruxelles: Musees Royaux d’Art et d’Histoire. Copyright © 1989 Michael Jansen. Reprinted with permission; 10.22, Redrawn with permission from original illustration by J.M Kenoyer; 10.26, From The Riddle of the Pyramids by Kurt Mendelssohn © 1974 by Kurt Mendelssohn. Thames &
Hudson, Ltd., London. Reprinted with permission; 10.28, From Michael Hoffman, “Where Nations Began,” Science, October 1983, pp. 42–51. 1983; 10.34, From The Riddle of the Pyramids by Kurt Mendelssohn © 1974 by Kurt Mendelssohn. Thames & Hudson, Ltd., London. Reprinted with permission; 10.35, After The Riddle of the Pyramids by Kurt Mendelssohn. Thames & Hudson, Ltd., London; 10.38, From Wenwu, 1984; 10.39, K.C. Chang, Studies of Shang Archaeology. Copyright © 1986 Yale University Press. Reprinted by permission of the publisher, Yale University Press; 10.43, From K.C. Chang, “Food & Vessels In Ancient China,” Translations of the NY Academy of Sciences, Vol. 35, No. 6, 1973. Copyright © 1973 NY Academy of Sciences. Used with permission; 10.46, © Hsien-Min Yang National Geographic Image Collection; 10.57, From Great Zimbabwe by Peter S. Garlake. Copyright © 1973 Thames & Hudson. Ltd., London. Reprinted with permission; 10.58, From Great Zimbabwe by Peter S. Garlake. Copyright © 1973 Thames & Hudson. Ltd., London. Reprinted with permission; 10.59, From Great Zimbabwe by Peter S. Garlake. Copyright © 1973 Thames & Hudson. Ltd., London. Reprinted with permission Chapter 11 11.12, Courtesy the Forschungsinstitut fur Alpine Vorzeit; 11.13, Courtesy the Forschungsinstitut fur Alpine Vorzeit; 11.15, Courtesy Dr. Ullrich Ruoff; Table 11.2 The Distribution of Animals During the Later Occupation at Charavines, Based on 504 Individuals. Centre de Documentation de la Prehistoire Alpine. Reprinted with permission from Aimé Bocquet; 11.16, Courtesy of A. Houot/Centre de Documentation de la Prehistoire Alpine. Reprinted with permission from Aimé Bocquet; 11.25, Christopher Chippendale, Stonehenge Complete, Illustration 240, p. 269. Copyright © Crown. NMR. Used with permission; 11.27, After Stonehenge Complete by Christopher Chippendale, Thames & Hudson, Ltd. Reprinted with permission; p. 571, Siegfried Sassoon, from “The Heart’s Journey: 1928.” With permission from Heinemann Publishers, Ltd.; 11.31, Past Worlds: The Times Atlas of Archaeology, edited by Chris Scarre. Reprinted by permission of HarperCollins Publishers Ltd. and Chris Scarre. © Chris Scarre 1988; 11.39, J. Jensen, 1979, Bronzealderen, Vol 1, Copenhagen: Sesam Publishers. Published by permission of the National Museum of Denmark; 11.42, Peter Vilhelm Glob, The Mound People; 11.44, J. Jensen, 1979, Bronzealderen, Vol 1, Copenhagen: Sesam Publishers. Published by permission of the National Museum of Denmark; 11.46, Reprinted from Prehistoric Europe by T. Champion et al., Copyright © 1984. With permission from Elsevier; 11.52, Wheeler’s interpretation of the construction phases at Maiden Castle. Copyright © Society of Antiquaries of London. Used with permission Chapter 12 Table 12.1, The Mission of UNESCO’s World Heritage Program. Credit: Mission Statement of UNESCO’s World Heritage Centre. Used with permission; Table 12.2, A Partial List of World Heritage Archaeological Sites. Credit: Partial list of UNESCO’s World Heritage List. Used with permission; Table 12.4, 8 ethical principles from website saa.org/aboutSAA/ethics.html. Courtesy of the Society of American Archaeology: www.saa.org.
_ Credits
C-5
pri31057_cre.ppC1-C6.qxd
_ _
C-6
Credits
10/2/09
8:12 AM
Page C-6
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
Page I-1
Index < Page numbers for illustrations and tables are italicized. Aborigines, 143, 143, 194 Abri Blanchard site, 141, 141 absolute dating method, 46, 48, 49 Abu Hureyra site, 204, 207, 212, 214–217, 215–217, 225–226, 226 accelerator mass spectrometry (AMS), 121, 146, 146, 227, 227, 241, 244, 249, 258–259 Acheulean handaxes, 79, 90, 98, 98, 99, 99–100 achieved status, 293 Acosta, Jorge, 342, 373 acropoli, 360 Adams, Douglas, 64 Adena complex, 269 adobe, 301 adzes, 181, 182, 274 Aegean Bronze Age, 532–534 Knossos site, 509, 535–537, 535–537 site map, 532 trade, 532, 533–534 writing, 533, 533, 534 aerial photography, 9, 9 Africa agriculture, 202, 263, 263, 492–494, 496–497 climate/environment, 492–493 site maps, 440, 492 state development, 439, 497–499 technology, 492 timeline, 440 trade, 439, 492, 493, 494–495, 495, 497, 499–500 See also South Africa; sub-Saharan Africa African Eve hypothesis, 115 Agamemnon, 507, 539, 540 age at death, 23, 192–193, 225, 225 See also life expectancy agriculture, 198–265 Africa, 202, 263, 263, 492–494, 496–497 breast-feeding, birth spacing and, 260–261, 264 China, 479, 483 climate/environment and, 205–206, 208, 211, 214, 215, 215 division of labor and, 28 East Asia, 201–202, 204 Egypt, 224, 224, 462–463, 471 food storage, 199, 201, 207, 222, 234, 283, 315, 407, 430–431 food surpluses, 207, 272, 315, 474, 502, 553 Indus civilization, 454 Mesoamerica, 204, 208, 211, 243–245, 250–251, 253, 262, 322–323, 363–364, 364 Mesopotamia, 448 Mexico, 202, 204, 243–245, 246–249, 259 Neolithic Europe, 507, 508, 508–509, 511–512, 522–523, 523, 527, 555 Neolithic period and, 204, 207, 208, 211–213, 261, 262–263 organization and, 28–29, 212, 264 origins of, 29, 162, 165, 196, 199–202, 204 primary domestication centers, 201–202, 203, 204, 208 regions and timeline of, 201, 201 secondary domestication centers, 201 sedentism and, 207, 226, 260, 262, 263, 264
site map and timeline, 203 South America, 202, 204, 254–255, 256, 392, 394, 395, 396, 398, 407, 411, 416, 417, 418, 430–431 Southeast Asia, 488, 491 Southwest Asia, 201–202, 202, 204, 205–206, 208, 211–213, 218, 224, 224–226, 226, 234, 261, 262, 441–443 spread of, 204, 262–264 as subsistence pattern, 27 See also chinampas; native North American agriculture; specific products/sites Ahkenaten, (Egyptian pharaoh), 471 Ahuitzotl (Aztec ruler), 383 ‘Ain Ghazal site, 207, 207, 227 ‘Ain Mallaha site, 204, 209–210, 209–210, 262 Alexander the Great, 472, 541, 557 Algaze, Guillermo, 446 alpacas, 202, 257 alphabets, 453 altiplano, 416 Alva, Walter, 391, 411–413 Amazonian South America, 392 Ambassador Monument (La Venta site), 341, 341 Ambrona site, 90, 90, 97 American Bottom, 283, 286 American Indian Religious Freedom Act, 187 Americas. See Archaic period; Mesoamerica; native North Americans; New World migration; South America Amratian tradition, 463, 464 AMS. See accelerator mass spectrometry analysis, 18–24, 33 artifacts, 18–20, 20 ecofacts, 18–22, 21–22 features, 19, 22, 22–23 settlement patterns, 24, 25, 25, 26 sites and settlements, 23–24, 24, 26 soil, 10, 17–18, 22 ancestor worship, 222, 420, 527, 568 Ancestral Pueblo culture, 270, 301, 307–308, 310 ancient DNA, 115, 115–116, 122, 122 Andersson, John Gunnar, 85 Angkor state, 439, 488–491, 488–491 Angkor Thom, 489, 489, 491 Angkor Wat, 489, 489, 490, 490 animals domestication of, 125, 161, 196, 200, 241 extinctions, 90, 90, 95–97, 95–97, 127, 148, 148, 150, 151–152, 151–153, 157, 165, 166, 268 specialists, 19 See also faunal remains; mammals; specific animals annealing, 408, 478 anthropogenic processes, 504 anthropological archaeology, 8 anthropomorphic figures, 232, 233, 402, 404 antiquities, import and sale of, 563, 566, 567, 570 Antiquities Act (1906), 567 An-yang site, 439, 476–478, 478 Ao site, 476, 477, 477 Arago skull, 89 archaeoastronomy, 426 archaeobotany, 19, 218–220, 218–220 Archaeological Conservancy, 564 archaeological heritage, 563–564
Archaeological Resources Protection Act (1979), 187, 567 archaeology basics of, 32–33 careers in, 572–573 defined, 1 ethical issues, 160, 569, 570, 571 fundamentals of, 7–14 legislation and, 567–569 lessons of, 574 ownership issues, 563, 564, 566 public information about, 570–572 responsible, 569–572 theories, 25, 27, 33, 106, 107, 107 value of, 561–562, 566 Archaeology Week programs, 570, 571 archaeometrists, 19 archaeozoology, 19, 21, 224–226 Archaic Homo sapiens, 89, 105 Archaic period, 165–167 Carrier Mills site, 167, 187–191, 187–191 native North Americans, 268, 269 Poverty Point site, 9, 269, 272–275, 272–275 architecture Abu Hureyra site, 214, 216, 217, 217 ‘Ain Mallaha site, 209–210, 262 Ban-po-ts’un site, 237, 237, 238, 475 Cahokia site, 284, 286, 474 Çatalhöyük site, 228–230, 229–230 Chaco Canyon site, 307–309, 307–310 Charavines site, 521–522, 522 Dolni Vestonice site, 128–129, 128–129 El Mirador site, 337–340, 338–339 Eridu site, 442–443, 444, 445, 445, 447 Great Zimbabwe site, 496, 497–500, 498–500 Huánuco Pampa site, 429–432, 430–432 Indus civilization, 454–456 Iroquoian, 295–299, 296–297 Jericho site, 222–223, 223 Jomon period, 184, 184–185 Khok Phanom Di site, 242 Maya civilization, 323 Mehrgarh site, 232, 232 Monte Albán site, 330, 342–346, 342–346, 354, 386 Monte Verde site, 154–156, 155–156 nonresidential, 329–330, 329–330 Ozette site, 313, 314, 314 Palenque site, 365–366, 365–368, 368 Paleolithic period, 262 Pincevent site, 139, 139 Poverty Point site, 274 San José Magote site, 327–328, 328, 329–330, 330, 342–343 San Lorenzo site, 333–334, 334 Snaketown site, 300–303, 301–302, 305, 305 Teotihuacan site, 329, 349–355, 350–353, 386 Tikal site, 329, 357–360, 357–360, 368 Tula site, 371–372, 371–372 See also monumental architecture archives, 8, 11, 13, 14 Ardipithecus ramidus, 40–41, 66 Arens, Williams, 385 argon-argon dating, 49 Arnold, Dean, 377 art Aborigine, 143, 143
Index
I-1
_
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
art (continued) Çatalhöyük site, 228–229, 231, 231 Celtic, 547, 547–548 Chichén Itzá site, 376, 377 Egyptian, 200, 200, 464 ideologies and, 31 Indus civilization, 456 Knossos site, 536–537 Paleolithic period and, 31, 79 South America, 392, 405, 414, 418 Southeastern Ceremonial Complex, 289–290, 290, 293 Teotihuacan site, 354, 354–355 See also petroglyphs; pottery; Upper Paleolithic art arthritis, 193 artifacts analysis of, 19–20, 20 context among, 15, 21 defined, 8 exchange of, 27, 28 lithic, 64 provenience of, 15 recording location of, 10 See also art; grave goods; technology ascribed status, 293, 295 Asia, 440 See also East Asia; Southeast Asia; Southwest Asia Asikli site, 231 Asperio site, 395–396, 396 assemblages, 64, 79 association dating method, 48 Atapuerca site, 89, 91–93, 91–94 Atauhualpa (Inca ruler), 423 Atlantean column, 372, 373 atlatl, 280, 283 Aubrey holes, 528, 529 augers, 10 Aurignacian period, 124, 136 aurochs, 96, 96, 170, 170 Austin, Alfredo López, 384 Australia, 107, 142–143, 142–143, 169, 194 australopithecines, 41, 42, 53, 66, 67 Australopithecus, 41 Australopithecus afarensis, 36, 41, 47 Australopithecus africanus, 42, 42 Australopithecus anamensis, 41 Aztec civilization, 267–268, 291, 321 codices, 198, 199, 204, 381, 381, 383, 384, 385 corn and, 198, 199, 204, 247, 385 European-Native American contact, 198, 199, 321, 349, 378, 380, 382, 384–385 marketing, 383, 383, 385, 433 monumental architecture, 320, 321 Sacred Almanac, 370, 370 Tenochtitlán site, 321, 356, 372, 378–380, 378–382, 382, 385, 387 Toltec civilization and, 371 writing, 369
_ _
baboons, 55, 98 Badarian sites, 463 Bahn, Paul, 17 bajos, 339 Balakot site, 460 ball courts Hohokam culture, 301–303, 302–303 Mesoamerican, 329, 356, 356, 375, 376, 376 Bambandyanalo site, 497 Ban Chiang site, 241, 242 bananas, 263, 497 Bandkeramik culture, 508 bands, 29, 315 bangles, shell, 460–461, 461 Ban-po-ts’un site, 204, 236–238, 236–238, 475 Bantu-speaking peoples, 496 barley, 208, 210, 212, 218, 219, 262
I-2
Index
Page I-2
barrows, 527, 544–546, 544–546 bartering, 281, 383, 555 Bar-Yosef, Ofer, 223 Basal Paleolithic, 59, 73, 79 baskets, 234, 255, 257, 295, 313, 313 Basu, Sudipta, 452 Beadle, George, 247 beans, 202, 244, 247–248, 256, 262, 322–323, 385 Bender, Barbara, 207 Bering Land Bridge, 149, 169, 268 Beringia, 147, 147–150 Bernal, Ignacio, 342 bier, 242 bifacial tools, 64 Big Bang, 2, 2 Bilzingsleben site, 90 binary-coded systems, 394 Binford, Lewis, 139, 206, 207 Bingham, Hiram, 426 bioarchaeologists, 19, 23 biological anthropology, 7–8 biological evolution, 5, 6, 6–7, 32 bioturbation, 14 bipedalism, 35, 40, 41, 51, 66, 68–69 Birth of the Gods and the Origins of Agriculture (Cauvin), 207 birth order, 30 birth spacing, breast-feeding and, 260–261, 264 Biskupin site, 509, 509 bison, 95, 95, 133, 134, 152, 158, 159, 167 Black, Davidson, 85–86 Black Earth site, 187–191, 187–191 blades, 27, 79, 79, 123–124, 123–125, 138–139 blanks, 124, 124, 125 Blanton, Richard, 342, 346 Bocquet, Amié, 520–521 body hair, 5, 67, 75 body size, 37, 66, 66–67, 192, 192–193 bog bodies, 509, 550–551, 550–551 bola stones, 155 bones chemistry, 176–177, 176–178, 182, 253 faunal analysis and, 21–22, 21–22 hyoid, 109 oracle, 475, 475, 476 samples, 17 skeletal evidence and, 192–193, 192–193 bonobos, 39 borers, 100 Borum Eshøj site, 509, 544–546, 544–546 Boserup, Esther, 502, 503 Boswell, Percy, 61 Boule, Marcellin, 117 Bourget, Steve, 409 bow and arrow, 125, 283 bow-drill, 233 Boxgrove site, xxiv, 1, 89 Boyd, Robert, 315 Braidwood, Robert, 206 Brain, C. K., 52, 53–54 brain organization, 109 brain size hominin evolution and, 35, 37, 42, 43, 51, 66–68, 66–68, 161 of Homo erectus, 66–67, 67, 74, 74, 87–88, 103 Neanderthals, 118 Brainerd, George, 363 Braudel, Fernand, 504 Braun, David, 187 braziers, 327 breast-feeding, 35, 67, 260–261, 264 breccia, 52–53 Brezillon, M., 138 bridges, 428, 428 Broca’s area, 110 Broken K Pueblo site, 305 bronze, 533, 534 Bronze Age Europe, 78
Minoan civilization, 533–534, 536–537, 555 Mycenaean civilization, 506, 506, 507, 509, 538–541, 538–541, 555 Northern Europe, 509, 542–546 political organization, 534, 541 technology, 542, 542–543, 555 trade, 542–543, 543 See also Aegean Bronze Age brow ridges, 118, 119 Brückner, E., 81 bullae, 451, 451–453 Burger, Richard L., 401 burials Abu Hureyra site, 217 ‘Ain Mallaha site, 209–210, 209–210 analysis of, 22, 23, 24 Australia, 143 Ban-po-ts’un site, 237–238 Borum Eshøj site, 544–546, 544–546 Bronze Age Europe, 538–540, 539–541 Cahokia site, 286, 286, 288 Carrier Mills site, 189–191, 189–191 Çatalhöyük site, 228–230, 230 China, 439, 477, 478, 481, 483–487 culture and, 23 Dolni Vestonice site, 129, 129 Egypt, 439, 456, 463–464, 467, 467, 469, 471, 473 Hopewell sites, 276–280 Indus civilization, 456 Iron Age Europe, 509, 548–549, 548–549 Jenné-jeno site, 494, 494 Jericho site, 222, 222 Jomon period, 184, 185 Khok Phanom Di site, 242, 242 Mehrgarh site, 232, 232–233 Mesolithic period, 164, 165, 166 Mesopotamia, 449 Monte Albán site, 345 Moundville site, 289, 291–292 Neanderthals, 120, 120, 122, 192 Palenque site, 366–368, 367–368 Paleolithic period, 79, 93 Poverty Point site, 275 Sipán site, 411–414, 412–414 Teotihuacan site, 351, 354–355 Varna site, 513, 513–515, 515 Vedbaek site, 164, 165, 166, 172–175, 173–174, 177 burins, 100 Butler, Brian, 187 Butzer, Karl, 472 Cabrera, Rubén, 351 cacao beans, 383 caches, 285, 286 Cactus Hill site, 149 Caen, Herb, 80 Cahokia site, 266, 266–267, 269, 283–286, 283–288, 288, 322, 474 Cahuachi site, 410 Calakmul site, 360, 362, 365, 366 Calendar Round, 370, 370 calendars, 126, 141, 341, 361–362, 369–370, 370, 451 See also season awareness calibrated dates, 146 Cambodia. See Angkor state camelids, 202, 257, 394 camels, 494 canals. See water management cannibalism, 94, 112, 120, 161, 384–385 Cannon-Brookes, Peter, 15 Caracol (Mayan lord), 362 Caral site, 395, 395, 396, 399 carbon dioxide (CO2), 145, 205–206, 206, 208 carbon isotope ratios, 176–177, 176–177, 182 Carlyle, Thomas, 100
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
Carnac site, 525, 525 Carneiro, Robert, 502, 503 carnelian, 448 Carrier Mills site, 167, 187–191, 187–191 Carter, Howard, 473 Caso, Alfonso, 342, 345 caste system, 28 Çatalhöyük site, 204, 227–231, 227–231 cats, 213 cattle, 213, 215, 217, 223, 233, 496 Cauvin, Jacques, 207 cave bears, 95, 95, 96, 161 cave paintings, 31, 96, 127, 130–135, 130–135, 182, 182 caves, 352 caymans, 401 Celtic civilization, 509, 557 See also Iron Age Europe celts, 278, 279, 293 cemeteries. See burials Cenabrum site, 557 cenotaphs, 514 cenotes, 374–375, 375, 377 Cenozoic era, 4, 5, 32, 35 ceramic specialists, 19 ceramics, 185, 300 See also pottery Cerro Azul site, 394 chac mools, 375, 377 Chaco Canyon site, 269, 270, 301, 306–309, 306–310 Chad site, 40 Chan Chan site, 393, 419–422, 420–422 Chanca people, 423 Chang, K. C., 475, 479, 480, 481 change, 5–7, 32, 574 Charavines site, 508, 520–523, 520–523 charcoal samples, 17, 145 charnel houses, 290, 291 Châtelperonnian period, 124 Chatters, James, 160 Chauvet cave site, 134 Chavín de Huántar site, 392, 400–403, 400–403, 404, 434 Chavín Horizon, 401, 403, 404 Cheops (Khufu) (Egyptian pharaoh), 469 Chephren (Khafre) (Egyptian pharaoh), 469 cherts, 274, 275 Chesterton, Gilbert K., 318 Chetro Ketl site, 307, 307 Chiang-chai site, 238 chicha, 431 Chichén Itzá site, 325, 329, 356, 374–377, 374–377 Chichimec people, 378 chiefdoms, 30, 315, 315–316, 318 Childe, V. Gordon, 205, 206, 501 chimpanzees comparison to humans, 37, 40, 40, 67, 109, 109, 115 hunting and, 55 technology and, 32, 32, 63, 68 Chimú civilization, 393, 419–422 China, 475–487 agriculture, 479, 483 An-yang site, 439, 476–478, 478 culinary traditions, 480–481, 480–481 Han dynasty, 480, 487 Han period, 480 monumental architecture, 476–477, 477 Qin dynasty, 439, 483–484, 487 Shang dynasty, 439, 475–479, 481 site map, 476 state development, 439, 475–479, 482–487, 503–504 status differentiation, 439, 475, 477–478, 481, 482 technology, 487
Page I-3
trade, 483 warfare, 479, 482, 483, 484 water management, 483, 485 writing, 451, 453, 475, 478–479, 484 Xia dynasty, 475–476 Xianyang, 439, 479, 482–487 Zhou dynasty, 475, 479, 480–481, 482–483 chinampas, 363, 379, 379 Chiripa site, 416, 417 chlorite, 458 Cholula site, 474 Christianity, 509, 558 Cieza de León, Pedro, 433 circumscription, 503 citadels, 534 ciudadelas, 419–422, 420–422 civilizations, characteristics of, 501 civilized societies, 438 Clactonian assemblages, 100 Clark, Desmond, 98 Clark, John, 351 class, 28, 30 See also status differentiation Classic period (Mesoamerica), 325 See also Maya civilization classical archaeologists, 19 classification, 20, 35, 39 clay tablets, 447, 448, 449, 451–453 cleavers, 100 climate/environment Africa, 492–493 agriculture and, 205–206, 208, 211, 214, 215, 215 hominin evolution and, 66, 73 Mesoamerica, 322 native North Americans and, 311, 313, 318 Neanderthals and, 118–119, 121–122 Pleistocene epoch, 81–83, 81–84, 89, 101, 103, 147–148, 151–153, 157 Postglacial period, 165, 169, 169–170 South America, 391, 399, 399, 416 clothing, 107, 126, 193, 480, 517, 517, 545, 545 Clovis points, 148, 148, 149–150, 150 CO2. See carbon dioxide coca, 256, 417 Codex Florentino, 198, 199, 204, 383 Codex Mendoza, 381, 381 codices, 198, 199, 204, 323, 381, 381, 383, 384, 385 Coe, Michael D., 332, 333 Coe, William, 359 coercive model of state development, 502 Coggins, Clemency, 362 Cohen, Mark, 206–207, 208 Columbus, Christopher, 147, 246, 267, 321, 391 comal, 344 community plan analysis, 305, 305 complexity, 1, 29, 30 conflicts. See warfare Confucius (Kongzi), 480, 483, 484, 562 Conkey, Margaret, 134 conquistadors, 292, 317, 322 Conrad, Geoffrey, 420 context, among artifacts, 14, 21 Conyers, Lawrence, 11 cooking, 101, 175, 213, 274, 305, 327, 344 See also culinary traditions, Chinese Copán site, 362, 365, 367 Coppergate site, 16, 16 coprolite, 398, 399 cord-marking, 184, 237 core stones, 63, 64, 64 corers, 10 coriander, 511 corn (maize), 246–249, 246–249, 262 Aztec civilization and, 198, 199, 204, 247, 385 in Mesoamerica, 322–323 native North Americans and, 258, 268, 269, 276, 280, 283, 295, 300
Cortés, Hernán, 321, 322, 380, 382, 385 corvée labor, 328 cosmologies, 30, 365–366 Cosquer cave site, 135, 520 Costantini, Lorenzo, 232 Costin, Cathy L., 460 cotton, 233, 396 counting systems, 126, 140–141, 140–141 Cowgill, George L., 355 Coxcatlán Cave site, 250, 250, 251 Coyolxauhqui (Aztec deity), 378 craft specialization, 29 Great Zimbabwe site, 499 Indus civilization, 233, 454 Mehrgarh site, 233 Mesopotamia and, 445, 447–448 trade and, 438, 460 cranial capacity. See brain size cranial deformation, 23 Creamer, Winifred, 395 Creek people, 292 cremations, 23, 143, 275, 276–277, 529 Crete, 509, 532, 533, 534, 535–537, 535–537 CRM. See cultural resource management Cro-Magnons, 124 crowns, Egyptian, 464, 465, 465, 468 Crumley, Carole L., 438 Cueva Blanca site, 243, 244 Cuicatlán Canada site, 345 Cuicuilco site, 349, 350 Culhuacan state, 378–379 culinary traditions, Chinese, 480–481, 480–481 cultigens, 166–167 cultivation, 200, 208 cults of the dead, 527 cultural anthropology, 7–8 cultural heritage, 564, 569 cultural resource management (CRM), 187, 573 culture agriculture and, 207, 208 burials and, 23 defined, 5, 32–33 development of, 5–6, 6 humans and, 37, 69 cuneiform, 448, 453 Curtis, Garniss, 48 Cusi Inca Yupanqui (Inca ruler), 423 cutmarks, 55, 55, 63, 224 Cuzco site, 393, 423–425, 424–425, 427 Cyclopean walls, 540 Cyprus, 213, 222 D’Altroy, Terence N., 433 Danta Complex, 338, 340 danzantes, 328, 343, 344, 346 Dart, Raymond, 53–54 Darwin, Charles, 6–7, 38, 117 dating, 48–49 absolute, 46, 48, 49 radiocarbon, 17, 48, 121, 145–146, 145–146 radiopotassium, 45, 48, 48–49, 50 stratigraphy and, 15–17, 48 table, 49 uranium-series, 93 datum point, 14 Daugherty, Richard, 312 De Hangen site, 179, 179, 180–181, 182 de Landa, Diego, 377 de Soto, Hernando, 292, 317 Deacon, Hilary, 112, 113 débitage, 63 Decker-Walters, Deena, 259 decoration. See art dendrochronology, 146, 308–309, 309, 521 Denevan, William M., 392 Denham, Tim, 263 dental enamel hypoplasia, 193 dental mutilation, 23
Index
I-3
_
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
Diamond, Jared, 262 Díaz, Bernal, 321 Dickhaut, John, 452 Diehl, Richard, 332, 373 diet. See nutrition Dikov, Nikolai, 148 Dillehay, Tom, 154 diseases, 23, 192, 193, 193, 260, 292, 299, 317 division of labor, 28, 102, 460 Dmanisi site, 75, 103 DNA (deoxyribonucleic acid), 114–116, 115–116, 122, 122 Doebley, John, 247–248 dogs, 125, 161, 196, 200, 241 dolmens, 525, 525, 526, 526 Dolni Vestonice site, 127–129, 127–129 domestication, 125, 161, 196, 199, 200–201, 220, 241 See also agriculture Donnan, Christopher, 408, 409, 413 Douglass, A. E., 308 Drachenhöhle site, 95 Drake, Robert, 48 Draper site, 269, 270, 295–298, 295–299 dryopithecines, 35 Durrington Walls site, 531
_ _
E Group, 340 Earle, Timothy, 401 Earth’s orbit, variations in, 83–84 East Asia agriculture, 201–202, 204, 239 Ban-po-ts’un site, 204, 236–238, 236–238, 475 Khok Phanom Di site, 204, 240–242, 241–242, 488 See also Southeast Asia Easter Island, 144, 563 ecofacts, 18–22, 21–22 economic specialization, 23–24, 447, 460–461 economy, 27, 28–29 edge hypothesis, for agriculture origins, 206 effigies, 275, 275, 278–280, 279, 282, 282, 290, 298, 298 egalitarian societies, 28, 30, 31 Egtved site, 544 Egypt, 462–472 agriculture, 224, 224, 462–463, 471 art, 200, 200 Hierakonpolis site, 439, 463–465, 467 Middle Kingdom, 470 monumental architecture, 439, 465 Napoleon and, 564 New Kingdom, 471, 492 Old Kingdom, 468–470, 473 political organization, 439, 462–474 pyramids, 439, 468–470, 468–470, 473–474 site map, 462 status differentiation in, 436, 436, 437 trade and, 468, 470 unification of, 465, 467, 468 water management, 462–463, 465, 471, 471, 472 writing, 451, 453 El Amra site, 463 El Baldari site, 463 El Gerza site, 464 El Mirador site, 325, 337–339, 337–340, 358 El Niño, 399 El Paraíso site, 396–398, 397–398, 399, 434 El Tajin site, 356, 356 Elands Bay Cave site, 166, 179–181, 179–182 Eldredge, Niles, 7 Elgin marbles, 564, 566, 566 Elias, J. N., 244 Eliot, T. S., 574 emblem glyphs, 361, 362, 365 Emerson, Thomas, 283 empires, 254, 321, 322 endocasts, 109–110, 110
I-4
Index
Page I-4
Engel, Frederic, 397 environmental impact studies, 187, 573 Environmental Protection Agency, U.S., 187, 573 epigraphy, 361 epiphyses, 193 epochs, 4 See also specific epochs equinoxes, 274, 285 eras, 4 See also specific eras Erickson, Clark L., 364 Eridu site, 438, 442, 442–444, 444, 445, 445, 447 Erlitou site, 475, 476 Eskimos, 118, 139, 150, 177, 194, 194 Essay on the Principle of Population (Malthus), 6 estrus, 68 estuaries, 322 ethics, 160, 569, 570, 571 ethnocentrism, 317, 323, 423 ethnographies, 25, 27, 323 ethnohistory, 268 Etiolles site, 138–139 Etowah site, 269 Europe, 506–558 Abri Blanchard site, 141, 141 Dolni Vestonice site, 127–129, 127–129 growth/evolution of civilization, 555–558, 556 Homo erectus, 89–93, 89–94, 101–103, 102 La Vache site, 140, 140 Lascaux site, 130–135, 130–135 medieval, 29, 29 Pincevent site, 138–139, 138–139 site maps, 89, 91, 510 timeline, 510 See also Aegean Bronze Age; Bronze Age Europe; European-Native American contact; Iron Age Europe; Neanderthals; Neolithic Europe; Postglacial Europe European-Native American contact, 269, 317–318 Aztec civilization, 198, 199, 321, 349, 378, 380, 382, 384–385 Inca civilization, 387, 391, 415, 423, 425, 426, 428 Evans, Arthur, 535 Evans, Clifford, 392 evolution biological, 5, 6, 6–7, 32 defined, 7 of life, 2–4, 3, 574 See also hominin evolution; Homo erectus; Homo sapiens evolution evolutionary tree, 39–41, 39–43 excavations, 1, 11–14, 12, 14, 33, 347, 569–570 defined, 8 director of, 11–13 field crew, 12 horizontal, or area, 17–18, 17–19 mapping, 8–9, 14, 14 selecting sites, 12, 14 test pits, 10, 15 vertical, 14, 15–16, 15–17 wet-site, 313 exchange, 27, 28, 222, 255, 281–282, 503 See also trade exotic materials, 27, 166, 216–217, 231, 282, 500 See also trade extraction sites, 24 extrasomatic means of survival, 69 facade, 340 facultative bipedalism, 40 Fagan, Brian, 434 Fairservis, Walter, Jr., 458 Falk, Dean, 110 family. See kinship systems farming. See agriculture
faunal remains Abu Hureyra site, 215–216, 225–226, 226 ‘Ain Mallaha site, 209–210 analysis of, 21–22, 21–22 Ban-po-ts’un site, 237 Carrier Mills site, 188 Çatalhöyük site, 229 De Hangen site, 182 European sites, 89, 90, 90, 94 Guitarrero Cave site, 257 Ho-mu-tu site, 239 Jericho site, 222–223 Khok Phanom Di site, 241 Klasies River Mouth Caves site, 112–113, 112–113 Lindenmeier site, 157–158 Mehrgarh site, 232–233 Monte Verde site, 155 Poverty Point site, 273 Sannai Marumaya site, 186 Vedbaek site, 173, 175 Zhoukoudian site, 86–87 features analysis of, 19, 22, 22–23 defined, 19 Feinman, Gary, 348 Feldhofer Cave site, 122 Feldman, Robert A., 396 Fertile Crescent, 206, 211, 212, 214, 218 fertility, 133, 137, 137, 260, 264 field crew, 12 field notes, 10, 12, 14 field schools, 572 field surveys, 8–10, 9–10, 12 field walking, 9, 9–10 fieldwork, 8–11, 33, 572 defined, 8 field surveys, 8–10, 9–10, 12 Finlayson, William, 296–297, 298 fire, use of Dolni Vestonice site, 128–129 humans and, 68 Kalambo Falls site, 98 Paleolithic period and, 79, 101, 161 Swartkrans site, 53, 54 “first family,” 45, 46–47 Fisher, Helen, 68 fishhooks, 313, 313 flake tools, 79, 100 See also Mousterian assemblages flakes, 63–64, 64 Flannery, Kent, 243, 245, 262, 328, 329, 334, 335, 351, 502 flint, 64, 79, 79 flintknapping, 64, 158, 158 FLK North site, 59 floodwater farming, 306, 307 flotation technique, 218–219, 219 fluted point, 149, 149 FMH. See fully modern humans Folsom points, 150, 158, 158–159 foragers, 165 See also hunter-gatherers Ford, James, 274 Formative (Preclassic) period (Mesoamerica), 325, 330 fossils defined, 37 hominin evolution and, 39–41, 52–53 Homo sapiens evolution and, 106 founder plants, 212 four quarters cosmology, 365–366 Fowler, Melvin, 285 Franchthi Cave site, 196, 507, 511–512, 511–512 friezes, 424, 566 fully modern humans (FMH), 5, 43, 73, 105 See also Homo sapiens evolution; Upper Paleolithic future shock, 574
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
Gainj people, 260–261 galena, 274, 275 Galindo site, 419 gallery grave, 525, 526, 526 Gamble’s Cave site, 61 Gargas site, 132, 132 G’a-to site, 483 gazelles, 209, 210, 214, 225–226, 232 Geertz, Clifford, 69 genetic evidence, 106, 107, 114–116, 114–116, 122, 122 genus, 39 geoarchaeologists, 19, 22 geoglyphs, 392–393, 410, 410 geological time, 3–5, 5, 32 geomorphic designs, 237, 238 geophysical prospecting, 10–11, 11 georadar (ground-penetrating radar; GPR), 10–11, 11 Gerzean period, 464–465 Gesher Benot Ya’Aqov site, 101 Ghab site, 215 Gheo-Shi site, 329, 329 giant beavers, 95, 95, 151 giant Irish deer, 96 giant sloth, 151, 152 Giza site, 439, 469–470, 469–470, 473, 473, 474 glacials, 81–82 glaciation, 81–84, 83, 89, 131, 147, 169 Glass Wreck site, 569 Glob, P. V., 550 glume, 220, 220 glyphs, 341, 345 goats, 213, 215, 217, 226, 232, 263 Gona River site, 63, 63 Good, Irene, 404 Goodall, Jane, 68 gorgets, 277 gorillas, 35, 40, 115 Gorman, Chester, 240–241 Gould, Stephen Jay, 7 gourds, 167, 186, 202, 259 government. See political organization GPR (ground-penetrating radar; georadar), 10–11, 11 Graham, Ian, 337 Gran Dolina site, 91, 92–94, 92–93, 103 grapes, 532 Grauballe Man, 550, 550 grave goods ‘Ain Mallaha site, 210 Ban-po-ts’un site, 238 Carrier Mills site, 190–191, 190–191 Çatalhöyük site, 228 defined, 190 Hopewell sites, 269, 276–277, 278–279 Khok Phanom Di site, 242, 242 Moundville site, 292, 293–294, 294 Poverty Point site, 275 status differentiation and, 293–294, 294, 407, 515, 546 Vedbaek site, 172 See also burials graves. See burials; grave goods; specific types Gravettian period, 124, 129 Great Sphinx, 469 Great Wall of China, 482, 482, 484, 487 Great Zimbabwe site, 439, 496–500, 496–500 Greece Early Iron Age, 509, 557 Elgin marbles, 564, 566, 566 Mycenae site, 506, 506, 507, 509, 538–541, 538–541 See also Aegean Bronze Age grids, 14 Griffin, James B., 289 Grotte de l’Hortus site, 120 ground-penetrating radar (georadar; GPR), 10–11, 11
Page I-5
Grove, David C., 334 guano, 397 Guilá Naquitz Cave site, 204, 243–245, 244–245, 247–249 guinea pigs, 202, 257, 394 Guitarrero Cave site, 204, 254–257, 254–257, 392 Gundestrup cauldron, 547 Haas, Jonathan, 395 habitual bipedalism, 40 Hadar site, 41, 44–47, 44–47, 48 Hadrian’s Wall, 509 Hadza people, 194 Halafian pottery, 442, 447 half-life, 49, 145 Hall, Robert L., 279–280 Hallam Çemi site, 213 Hallstatt period, 547, 548, 549 hammerstone, 63, 64 Han dynasty (China), 480, 487 Han period (China), 480 handaxes, 73, 100, 161 See also Acheulean handaxes handedness, 64–65 Hantman, Jeffrey L., 318 Harappa site, 232, 438, 455, 456 Harappan society. See Indus civilization hard-hammer technique, 100 Harlan, Jack, 218, 260 Harner, Michael, 385 Harris lines, 193 Hartley, L. P., 119 harvesting, 199 Hassan, Fekri, 439 Hastorf, Christine, 416 Haury, Emil, 301 Hawaii, 144 Hawass, Zahi, 474 health, 23, 193 See also diseases Hecht, Gary, 452 Helbaek, Hans, 219 hematite, 274, 275 hemp, 237 henge monuments, 524, 525 Hesse, Brian, 225 Hierakonpolis site, 439, 463–465, 467 hierarchical societies, 29, 29–30, 30, 31, 207, 264, 361 See also political organization; status differentiation hieroglyphs, 341, 368, 369, 369, 453, 465, 466 Higham, Charles, 241 Hill, Amos, 287 Hill, James, 305 Hillman, Gordon, 215, 218–219 Hinduism, 459, 490 historical archaeology, 8, 19 historical societies, 8 Hoabinhian complex, 240–241 Hodder, Ian, 228 Hoffman, Michael, 464, 467 Hohokam culture, 270, 300–304, 302–304 Holata Outina (Mississippian chief), 315, 315 Holloway, Ralph, 110 Holocene (Recent) epoch, 4–5, 84, 262 See also Postglacial period Homer, 533, 538, 539, 541, 557 hominids, 39–40 hominin evolution, 35–70, 562 bipedalism and, 35, 40, 41, 51, 66, 68–69 brain size and, 35, 37, 42, 43, 51, 66–68, 66–68, 161 brain v. body size, 37, 66, 66–67 climate/environment and, 66, 73 dryopithecines, 35 evolutionary tree, 39–41, 39–43 Hadar site, 41, 44–47, 44–47, 48 Kenyapithecus platyops, 36, 36, 40, 66
Laetoli site, 34, 35, 41, 48, 50–51, 50–51, 62 Leakey family, 61, 61–62 overview diagram, 41 overview table, 43 site map, 38 Swartkrans site, 52–53, 52–54 technology and, 42, 55, 63–64, 63–65 timeline, 38 See also Homo erectus; Olduvai Gorge site hominins, 40 See also hominin evolution Hominoidea, 35 hominoids, defined, 39 Homo antecessor, 94 Homo erectus, 73–103 Acheulean handaxes and, 79, 90, 98, 98, 99, 99–100 Atapuerca site, 89, 91–93, 91–94 brain size of, 66–67, 67, 74, 74, 87–88, 103 characteristics of, 74 European sites, 89–93, 89–94, 101–103, 102 evolution and, 41, 66–67, 87–88, 105 Hadar site, 44 Homo heidelbergensis and, 43, 89, 90, 93 Homo sapiens and, 105, 107 on Java, 75, 117 Kalambo Falls and Olorgesailie sites, 98, 98, 103 kinship systems, 102–103 migrations of, 89, 101, 103, 106 Nariokotome boy, 74, 74 nutrition and, 87, 98, 101–102 Pleistocene epoch and, 73, 81 reconstructions of, 75, 92 site discoveries, 74–76 site map and timeline, 77 skull, 36 Swartkrans site, 53, 54 Zhoukoudian site, 87–88, 103 Homo ergaster, 73, 74 Homo habilis, 42, 44, 61, 66, 110 Homo heidelbergensis, 43, 89, 90, 93 Homo neanderthalensis, 67, 90, 105, 117 Homo sapiens, Archaic, 89, 105 Homo sapiens evolution, 36, 41, 43, 67, 75, 105–162 Australia, 107, 142–143, 142–143 cave paintings, 96, 127, 130–135, 130–135 Dolni Vestonice site, 127–129, 127–129 genetic evidence and, 106, 107, 114–116, 114–116, 122, 122 Homo erectus and, 105, 107 Klasies River Mouth Caves site, 111–113, 111–113 Lake Mungo site, 142–143 language origins, 109–110, 109–110, 161 Lascaux site, 130–135, 130–135 Out of Africa/Multiregional theories, 106, 107, 107 Pacific Islands, 144, 144 Pincevent site, 138–139, 138–139 site maps, 108, 144 theory map, 107 timelines, 106, 108 See also Neanderthals; New World migration; Upper Paleolithic Homo sapiens neanderthalensis, 105 Homo sapiens sapiens, 36, 105, 112 See also fully modern humans Ho-mu-tu site, 236, 239 Hopewell, M. C., 269 Hopewell Interaction Sphere, 278, 280, 283 Hopewell sites, 269, 276–279, 276–280, 282 horizons, 335 See also Chavín Horizon; Olmec Horizon horizontal, or area excavations, 17–18, 17–19 Horus (Egyptian god), 463, 465, 466 household archaeology, 323 Hsi-pei-kang site, 477
Index
I-5
_
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
huacas, 405–406, 405–407, 409 Huánuco Pampa site, 394, 429–432, 430–432 Huascar (Inca ruler), 423 Huitzilopochtli (Aztec war god), 378, 379, 382 Huleh site, 215 human(s) characteristics of, 36–37, 37, 40, 50, 67–69, 74, 75 chimpanzees comparison to, 37, 40, 40, 67, 109, 109, 115 gender differences between apes and, 67–68 See also fully modern humans; Homo sapiens human behavior, 105, 161 human origins, 62, 562 See also hominin evolution; Homo erectus; Homo sapiens evolution human sacrifice autosacrifice, 327, 351, 384, 385 Mesoamerica and, 356, 376, 377, 379, 384–385, 384–385 native North Americans and, 286, 290, 293 San José Magote site, 327, 327, 328 hunter-gatherers, 562 Africa, 55, 206, 492–493 agriculture v., 199, 263 Australia, 143 contemporary, 194–195, 194–195 defined, 165, 166 in Egypt, 463 leisure time and, 206, 503 native North Americans as, 268, 270, 274, 276, 300, 311 organization, 28, 30 Paleolithic period, 79, 103, 133–134, 161–162, 211–212 Postglacial period, 165–167, 179–182, 183–186 prey of, 55 Southeast Asia, 488 subsistence pattern of, 27, 165, 196 hunting, 225–226 animal extinctions and, 127, 151–152, 157, 165, 166 dogs and, 125, 161, 196, 200 Dolni Vestonice site, 127–129 faunal analysis and, 21–22 Lindenmeier site, 157–159 Olduvai Gorge site, 55, 59 Pincevent site, 138–139 reindeer hunters, 138–139, 138–139 scavenging v., 55 during Stone Age, 112–113 Huron people, 297 Hurst-Thomas, David, 33 husbandry, 199–200 hyoid bone, 109 hypoglossal nerve, 110
_ _
Ice Age. See Pleistocene epoch Iceman (Ötzi), 508, 516–518, 516–519 iconography, 292 See also art ideographs, 369, 451, 452, 452, 453 ideologies, 30–31, 385 Iltis, Hugh, 247 in situ (“in place”), 11, 15 Inca civilization, 254, 267–268, 501 Chimú civilization and, 421–422 Cuzco site, 393, 423–425, 424–425, 427 European contact, 387, 391, 415, 423, 425, 426, 428 Huánuco Pampa site, 394, 429–432, 430–432 Machu Picchu site, 394, 425–426, 426 origin myth, 415 political organization, 391, 393–394, 418, 423–425, 429, 431–432, 433–434 road system, 394, 427–428, 427–428, 433 site maps, 393, 423, 427 status differentiation, 433–434
I-6
Index
Page I-6
timeline, 393 Tiwanaku site and, 415 warfare, 423, 429 incensario, 344 incest taboo, 32–33, 103 India. See Indus civilization indigenous (pristine) states, 322 Indus civilization, 454–461 agriculture, 454 craft specialization, 233, 454 Harappa site, 232, 438, 455, 456 Mohenjo-daro site, 232, 438, 455–457, 455–458, 460 shellworking in, 460–461, 461 site map, 454 state development, 438 trade, 458, 459 writing, 458, 459 inflorescences, 246, 247 intaglio, 456 integrative model of state development, 502 interglacials, 81, 82, 148, 151 interpretation, 24–31, 33 Inti (Inca god), 423, 424 intihuatana, 426 Iron Age Europe, 78, 556 Biskupin site, 509, 509 bog bodies, 509, 550–551, 550–551 Maiden Castle site, 509, 552–554, 552–554 Roman empire and, 547, 552, 554, 557–558 Vix site, 509, 548–549, 548–549, 552 Iroquoian culture, 269, 270, 295–298, 295–299 irrigation. See water management Isernia site, 89, 103 Islam, 495 isotopes, 48, 49, 145–146, 519 isotopic dating techniques, 48–49, 145–146 Itzcoatl (Aztec ruler), 371 Jacobsen, Thomas, 511 Jalieza site, 348 Japan, 166, 183–186, 183–186, 234, 235 Jarmo site, 206 Jarrige, Jean-François, 232 Jarrot, Nicholas, 287 Jasaw Chan K’awaiil I (Mayan ruler), 359, 360, 362 jasper, 233, 274, 275 Java, 58, 75, 117, 241 Jayavarman II (Khmer ruler), 489 Jayavarman VII (Khmer ruler), 491 Jefferson, Thomas, 267, 317 Jeffries, Richard, 187, 188 Jenkins, Dennis, 149 Jenné-jeno site, 439, 493–494, 493–495 Jericho site, 204, 207, 212, 221–223, 221–223, 227 jet, 310 Jiménez Moreno, Wigberto, 371 Joffroy, René, 548 Johanson, Donald, 42, 45–46, 46 Jolly, Clifford, 69 Jomon period, 166, 183–186, 183–186, 234, 235 jujubes, 239 Julius Caesar, 509, 547, 548, 557 Kalambo Falls site, 98, 98, 103 Kan Joy Chitam II (Mayan ruler), 367 Karanga culture, 497, 500 Karanovo site, 508 Kebara Cave site, 109, 121 Keeley, Lawrence, 65 Keillor, Garrison, 563 Kennewick Man, 160, 160, 568, 568–569 Kenoyer, J. Mark, 460 Kenyapithecus platyops, 36, 36, 40, 66 Kenyon, Dame Kathleen, 222–223 Khafre (Chephren) (Egyptian pharaoh), 469 Khamerernebti II, 470 Khmer civilization, 439, 488–491
Khok Phanom Di site, 204, 240–242, 241–242, 488 Khufu (Cheops) (Egyptian pharaoh), 469 Kibish site, 105 kill sites, 16, 152 kilns, 237, 238 Kilombe site, 99–100 king lists, 371 kingship. See political organization K’inich Janaab’ Pakal I (Mayan ruler), 366, 387 K’inich Kan Balam II (Mayan ruler), 367 kinship systems, 194, 501 defined, 28 hominin evolution and, 67 Homo erectus, 102–103 Kirchhoff, Paul, 356 kivas, 308–309, 308–309 Klasies River Mouth Caves site, 111–113, 111–113 Klein, Richard, 39, 107, 112 Knossos site, 509, 535–537, 535–537 Kojan, David, 417 Kolata, Alan, 417 Konso-Gardula site, 99 Koobi Fora site, 48, 53, 55, 62, 64, 65 Kowalewski, Stephen, 348 Krapina cave site, 120 kraters, 549, 549 Kua people, 195, 195 Kuk Swamp site, 263 Kunz Axe, 336 Kush, kingdom of, 492 Kusimba, Chapurukha, 499 Kutzbach, John, 83 La Chapelle-aux-Saints site, 117, 117 La Ferrassie site, 120 La Tène period, 547, 548, 557 La Vache site, 140, 140 La Venta site, 331–334, 332–333, 341, 341 lacquerware, 186 lactational amenorrhea, 260 Laetoli site, 34, 35, 41, 48, 50–51, 50–51, 62 Lagar Velho site, 122 lagunas, 333, 334 Lake Chad basin site, 40 lake dwellings, 508, 520 Lake Mungo site, 142–143 Lake Turkana site, 41, 62, 74 languages Mayan, 376 of native North Americans, 267 origins, 109–110, 109–110, 161 Lanzón (Great Image), 402, 403 lapidary, 274, 275 lapis lazuli, 233 Lapita pottery, 144, 144 Lascaux site, 130–135, 130–135 Late Stone Age, 111, 113, 166 lateralization, 109–110 Leacock, Eleanor B., 270 Leakey, Louis, 57–58, 61, 61–62 Leakey, Louise, 62 Leakey, Mary, 50, 50–51, 57–58, 61–62, 69, 100 Leakey, Meave, 40, 62 Leakey, Richard, 42, 44, 62, 74 Leakey family, 60–61, 61–62 Lee, Richard, 502, 503 leguminous plants, 177, 212 Lehner, Mark, 474 leisure time, 206, 502–503 lentils, 511 leopards, 52–53, 53 Leroi-Gourhan, A., 138 Les Trois Frères site, 132, 132 Levallois technique, 119 Levant, 209, 211, 213, 217, 231 Lewis-Williams, David, 229 Lhuiller, Albert Ruz, 366
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
Li Liu, 476 Libby, Willard, 145 lice, 107 life, evolution of, 2–4, 3 life expectancy, 67, 137, 190 See also age at death Lindenmeier site, 150, 157–158, 157–159 Lindlow Man, 550 lineages, 28, 354, 434 Linearband-keramik settlements, 26 Linnaeus, Carolus, 39 lintels, 359 lithic artifacts, 64 lithic specialists, 19 llamas, 202, 257, 394, 427 Lockyer, Norman, 531 Locmariaquer site, 525 locomotion, 41 loess, 127, 127 Loma de la Coyotera site, 345, 345 lomas, 395 Long Count calendar, 341, 361–362, 369–370, 370 Longgupo site, 75 longhouses, 270, 295–298, 296–298 Longshan period (China), 475, 476 looting, 15, 391, 411–413, 563, 566, 567 lost wax casting, 408 Lothal site, 460 Lovejoy, Owen, 69 Lower Paleolithic, 73, 78, 79, 90, 97, 100–103 Lo-yang site, 476 Lubbock, John, 78 Lucy, 42, 45, 46–47, 47 Luján, Leonardo López, 384 Lynch, Mark, 188 Lynch, Thomas F., 254–255 macaws, 310 maceheads, 464, 464, 465 Machu Picchu site, 394, 425–426, 426 MacNeish, Richard, 243, 248, 250–251, 253 macrofossils, 21 Magdalenian period, 124, 132, 137, 138 magnetite, 327 magnetometers, 11 maguey plants (agave), 244, 245, 245, 372 Maiden Castle site, 509, 552–554, 552–554 maintenance, of animals, 199 maize. See corn Makah Indians, 312 makers’ marks, 405–406, 406 male provisioning, 69 Mal’ta site, 129, 129 Malthus, Thomas, 6, 502 mammals, 35 in Pleistocene epoch, 95–97, 95–97, 151–152, 151–152 Postglacial period, 170, 170 See also woolly mammoths mammoths. See woolly mammoths Manhattan Project, 145 Mann, Charles C., 387 Mannion, Antoinette, 264 Mano Capac (Inca ruler), 424 manos, 300, 300 Manzanilla, Linda, 352 mapping, 8–9, 14, 14 Mapungubwe site, 497, 500 Marajó Island site, 392 Marcus, Joyce, 328, 329, 341, 345, 358, 394 marine resources Homo erectus, 101 Ozette site, 311–314 Postglacial period, 171, 171, 177, 177, 179–182, 183, 183–185, 196 South America, 395, 398, 399, 407 Maritime Hypothesis, 399, 399 marketing, 281–282, 383, 383, 385, 433 Marquesas, 144
Page I-7
marriage systems, 28, 468, 471 Marshack, Alexander, 140–141 Marshall, John, 455 Martin, Paul, 305 Marx, Karl, 504 mass spectrometer, 177 Massilia, 549 Matheny, Ray, 337 mating systems, 41, 46, 47, 102–103 Mauer skull, 89 Maxwell, Thomas, 291 Maya Blue pigment, 377 Maya civilization architecture, 323 Chichén Itzá site, 325, 329, 356, 374–377, 374–377 collapse of, 362, 367, 374 El Mirador site, 325, 337–339, 337–340, 358 Palenque site, 325, 362, 365–368, 365–368 political organization, 439, 488 texts, 323, 369 Tikal site, 325, 339, 354, 357–362, 357–362, 368 writing, 325, 341, 369–370, 369–370, 386 Mbuti Pygmy people, 194 Mc. Adams, Robert, 433 McHenry, Henry, 41 McIntosh, Roderick, 493–494 McIntosh, Susan, 493–494 Meadow, Richard, 233 megaliths, 524–530, 524–531, 555 Meggers, Betty, 392 Mehrgarh site, 204, 232–233, 232–233, 454 Mellaart, James, 227 Memphis, 439, 468, 469, 470, 471 Mendelssohn, Kurt, 473 Meneses, Susana, 391 menhirs, 524–525, 525 Menkaure (Mycerinus), 469, 470, 470 Merimde site, 463 Meroë site, 492 Mesa Verde site, 310, 560, 560, 561 Mesoamerica, 320–388 agriculture, 204, 208, 211, 243–245, 250–251, 253, 262, 322–323, 363–364, 364 ball courts, 329, 356, 356, 375, 376, 376 calendars, 341, 361–362, 369–370, 370, 451 carved stone importance, 341, 341 climate/environment, 322 corn in, 322–323 defined, 250 Guilá Naquitz Cave site, 204, 243–245, 244–245, 247–249 human sacrifice and, 356, 376, 377, 379, 384–385, 384–385 legacy of past, 386–387 Monte Albán site, 325, 328, 330, 342–346, 342–346, 354, 356, 386, 503 Olmec culture, 31, 325, 327, 332–336, 335–336, 337, 355 political organization, 361–362 pyramids, 287, 329, 332, 338, 338, 339–340, 474, 474 settlement patterns, 321, 347–348, 348 site maps, 243, 324 South America compared to, 392, 394, 433–434 state development, 321–323, 325 status differentiation, 382, 433, 434 Tehuacán sites, 204, 243, 247–249, 250–253, 250–253 timeline, 324 Toltec civilization, 371–373, 371–373, 375–376 trade, 354, 372, 382, 386 water management, 350, 363–364, 382 See also Aztec civilization; Maya civilization; Mexico; Teotihuacan site Mesolithic period, 12, 19, 32, 165–167, 222, 511, 520 defined, 166 See also Postglacial period
Mesopotamia agriculture, 448 craft specialization and, 445, 447–448 defined, 438 Eridu site, 438, 442, 442–444, 444, 445, 445, 447 site map, 441 state development, 438–439, 465 status differentiation, 444, 449 temple institution, 438, 443, 445, 449 Uruk site, 438, 444, 446, 446–449, 449, 451 warfare, 448–450 water management, 441–442, 443, 446 writing, 448, 451–452, 451–453 See also Southwest Asia Mesozoic era, 4, 5 mesquite trees, 244 metal detectors, 11 metallurgy, 242, 404 China, 476, 478, 479, 483 Egypt, 463 Neolithic Europe, 513–515, 514, 517 sub-Saharan Africa, 493 See also Bronze Age Europe; Iron Age Europe metates, 300, 300 meteor, extinctions from, 4, 153 Mexica people, 378 See also Aztec civilization Mexico agriculture, 202, 204, 243–245, 246–249, 259 Guilá Naquitz Cave site, 204, 243–245, 244–245, 247–249 pottery, 234–235, 234–235 Tehuacán sites, 204, 243, 247–249, 250–253, 250–253 Meyjes, Menno, 25 mica, 278, 279 microbands, 243, 251 micromorphologists, 19 middens, 188, 188 See also shell middens Middle Awash site, 40 Middle East, defined, 204 Middle Kingdom (Egypt), 470 Middle Paleolithic, 78, 79, 93, 105, 119 Middle Pleistocene, 76 Middle Stone Age, 111–113, 165 See also Mesolithic period migrations of Homo erectus, 89, 101, 103, 106 of Homo sapiens, 107 See also New World migration Milankovitch, Milutin, 83 Milankovitch forcing, 84 millennia, 507–509 millet, 186, 202, 261, 262, 263, 479 Millon, René, 350, 351, 355 Minoan civilization, 533–534, 536–537, 555 Miocene epoch, 4, 32 Mirabad site, 215 Mississippian culture Cahokia site, 266, 266–267, 269, 283–286, 283–288, 288, 322, 474 defined, 269 Moundville site, 269, 286, 289–292, 290–292, 293–294, 294 site map, 290 mit’a system, 406, 407, 420, 425 mitmaq, 429 mitochondrial DNA, 115–116, 115–116 Mixtec civilization, writing, 369 Moche civilization, 392–393 pottery, 406–408, 408 Sipán site, 390, 391, 393, 411–414, 411–414 Moche Sacrifice Ceremony, 409, 409, 413, 414 Mogollon culture, 270, 301, 305 Mohenjo-daro site, 232, 438, 455–457, 455–458, 460 molecular clock, 114
Index
I-7
_
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
Monks Mound site, 266, 266–267, 284–285, 285, 287–288, 288, 474 monochrome pottery, 442, 443, 443 Monos Complex, 338, 339–340 Mont Lassois, 509, 548–549, 552 Montague, Ashley, 50 montaña zone, 256 Monte Albán site, 325, 328, 342–346, 342–346, 503 architecture, 330, 342–346, 342–346, 354, 386 ball court, 356 Monte Verde site, 149, 154–156, 154–156, 254 monumental architecture Angkor state, 489–491, 489–491 Aztec civilization, 320, 321 Chan Chan site, 420, 421 Chichén Itzá site, 329, 374–376, 374–377 China, 476–477, 477 Cuzco site, 424–425, 424–425 defined, 287 El Mirador site, 338, 338–339 Hopewell sites, 278 Mesopotamia, 444, 445 Monks Mound site, 266, 266–267, 284–285, 285, 287–288, 288, 474 Palenque site, 365–366, 366–368, 368 Poverty Point site, 269, 273, 275 San José Magote site, 329 South America, 395–397, 401–403, 401–403, 434 Tenochtitlán site, 378, 378–379, 382, 382, 387 Tikal site, 358–360, 358–360 Tiwanaku site, 416–418, 417–418 Uruk site, 446–447 See also pyramids monuments, museums and, 571–572 Moore, A. M. T., 214 Moore, Clarence B., 291 Moore, Jerry, 287 Moorehead, Warren, 284 moraines, 82 Morley, Sylvanus, 363 Morris, Craig, 429, 430, 431 Morrison, Philip, 573 mortars, 209 Mortillet, Gabriel de, 78 Moseley, Michael, 398, 399, 405–406, 420 motifs, 282, 289–290, 291, 314 Moundville site, 269, 286, 289–292, 290–292, 293–294, 294 Mount Carmel site, 122 Mount Li, 484, 485 mountain, defined, 445 Mousterian assemblages, 79, 119, 119, 121 Movius, Hallam, 100 Multiregional theory, 106, 107, 107 multivallate defenses, 554 mural art, 131, 341, 354, 376, 377 See also cave paintings Murra, John, 427, 428, 432 museums, monuments and, 571–572 music, 128 m.y.a. (millions of years ago), 4 Mycenae site, 506, 506, 507, 509, 538–541, 538–541 Mycenaean civilization, 506, 506, 507, 509, 538–541, 538–541, 555 Mycerinus (Menkaure), 469, 470, 470
_ _
NAGPRA. See Native American Graves Protection and Repatriation Act Nakbe site, 337, 340 nanodiamonds, 153 Napoleon, Egypt and, 564 Naqada site, 463 Nariokotome boy, 74, 74 Narmer (Egyptian pharaoh), 465, 466, 467, 468 Native American Graves Protection and Repatriation Act (NAGPRA) (1990), 567, 567–568
I-8
Index
Page I-8
native North American agriculture, 202, 204, 258–259, 258–259, 270 Cahokia site, 283, 284 corn and, 258, 268, 269, 276, 280, 283, 295, 300 Draper site, 295 Hopewell sites, 276, 280, 282 Moundville site, 289, 291 Poverty Point site, 272, 274 Snaketown site, 300 native North Americans, 266–319 Archaic period, 268, 269 Cahokia site, 266, 266–267, 269, 283–286, 283–288, 288, 322, 474 Chaco Canyon site, 269, 270, 301, 306–309, 306–310 changing perspectives on, 270, 317–318 chiefdoms, 315, 315–316, 318 climate/environment and, 311, 313, 318 diversity of, 267, 301, 318 Draper site, 269, 270, 295–298, 295–299 Hopewell sites, 269, 276–279, 276–280, 282 as hunter-gatherers, 268, 270, 274, 276, 300, 311 Kennewick Man, 160, 160, 568, 568–569 languages of, 267 legislation for, 567, 567–569 Moundville site, 269, 286, 289–292, 290–292, 293–294, 294 Ozette site, 269, 270, 311–314, 311–314 Paleoindian period, 148–149, 148–150, 152, 156, 157–158, 166 political organization, 268, 286, 289, 318 pottery, 272, 272, 274, 274, 278–279, 286, 286, 289–290, 290, 302, 303, 303, 308, 308 Poverty Point site, 9, 269, 272–275, 272–275, 325 site maps, 271, 277, 290, 306 Snaketown site, 269, 270, 300–305, 300–305 timeline, 271 trade, 278–280, 281–282, 289–290, 299, 302, 310, 316 warfare, 292, 299, 315 See also Mississippian culture Natufian period, 209, 212, 214–215 natural habitat hypothesis, for agriculture origins, 206 natural heritage, 564 Natural Historic Preservation Act (1966), 567 natural selection, 5, 6–7, 32 Navajo people, 307 Nazca geoglyphs, 392–393, 410, 410 Neander Valley site, 107 Neanderthals, 25, 105, 117–120 burials, 120, 120, 122, 192 characteristics of, 117–118, 118 climate/environment and, 118–119, 121–122 comparisons with FMH, 118, 119, 119 distribution map, 120 fate of, 121–122 FMH and, 106, 121, 121–122, 124 La Chapelle-aux-Saints site, 117, 117 reconstructions of, 106, 118, 120 site and timeline map, 108 Near East, defined, 204 necropolis, 467 Neitzel, Jill, 310 Nelson Bay Cave site, 113, 113 Neolithic Europe, 507–509, 511–523 agriculture, 507, 508, 508–509, 511–512, 522–523, 523, 527, 555 Charavines site, 508, 520–523, 520–523 Franchthi Cave site, 507, 511–512, 511–512 Iceman, 508, 516–518, 516–519 megaliths, 524–530, 524–531, 555 technology, 511, 517, 523, 523 Varna site, 508, 513–515, 513–515 Neolithic period, 32, 177 Abu Hureyra site, 204, 207, 212, 214–217, 215–217, 225–226, 226
agriculture and, 204, 207, 208, 211–213, 261, 262–263 Ban-po-ts’un site, 204, 236–238, 236–238, 475 Çatalhöyük site, 204, 227–231, 227–231 defined, 78, 212 Jericho site, 204, 207, 212, 221–223, 221–223, 227 Mehrgarh site, 204, 232–233, 232–233 site maps, 212, 224, 236 nets, 245, 245, 396, 396 net-sinkers, 209, 210 New Guinea, 107, 142, 144, 169, 204, 260–261, 263 New Kingdom (Egypt), 471, 492 New World migration, 115–116 Beringia, 147, 147–150 Clovis points, 148, 148, 149–150, 150 Kennewick site, 160, 160, 568, 568–569 Lindenmeier site, 150, 157–158, 157–159 maps of, 116, 147 Monte Verde site, 149, 154–156, 154–156, 254 New Zealand, 144 Nittano site, 184–185 Nok culture, 493 nomes, 467, 468 Non Nok Tha site, 241, 242 nonlocal materials. See exotic materials; trade nonresidential architecture, 329–330, 329–330 Norse game piece, 557 North America. See native North Americans Northwest coast. See Pacific Northwest cultures nucleotide substitutions, 114, 114 numbers, 140–141, 140–141, 369–370, 369–370, 394 nutrition bone chemistry and, 176–177, 176–178, 182, 253 of contemporary hunter-gatherers, 194, 194 culinary traditions, Chinese, 480–481, 480–481 faunal analysis and, 21 hominin evolution and, 41, 55, 67, 69 Homo erectus and, 87, 98, 101–102 Kennewick man, 160 Monte Verde site, 156, 254 Neanderthals, 119 Postglacial period and, 165–167, 170, 173, 183–184, 186, 188, 196 skeletal analysis and, 23, 192, 193, 193 teeth and, 210, 260, 519 See also agriculture; cannibalism; marine resources; subsistence patterns Oase cave site, 122 oasis hypothesis, for agriculture origins, 205 obsidian, 159, 229, 231, 327, 350–351, 372 obsidian blades, 27, 281, 282, 373 obsidian core, 27 oca, 400 occipital bun, 118, 119 Old Copper culture, 167, 167 Old Kingdom (Egypt), 468–470, 473 Oldowan pebble tools, 59, 59, 64, 79, 100 Olduvai Gorge site, 56–60, 56–60 fossils at, 57–60, 57–60 Homo habilis, 42, 61 hunting, 55, 59 Leakey family, 57–58, 60–61, 61–62, 100 schematic cross section, 57 technology, 57, 57–60, 59–60, 64 Zinjanthropus, 58, 62 olives, 532 Olmec culture, 325, 327, 332–336, 355 defined, 332 ideologies in, 31 Olmec Horizon, 327, 333, 335–336, 335–336, 337, 392, 403 Olorgesailie site, 98, 98, 103
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
On the Origin of Species by Means of Natural Selection (Darwin), 6–7, 117 O’Neill, Eugene, 561 oppida, 509, 557 oracle bones, 475, 475, 476 organic materials, 13, 13, 20, 145–146 organization, 23, 27–29, 212, 264 See also kinship systems; political organization Orrorin tugenensis, 40–41, 66 Out of Africa theory, 106, 107, 107 ownership issues, 563, 564, 566 oxygen isotope ratios, 81–82, 81–82, 112 Ozette site, 269, 270, 311–314, 311–314 Pachakuti (Inca ruler), 415, 423, 424, 425, 426, 429 Pacific Islands, 144, 144, 263 Pacific Northwest cultures, 25, 269, 270, 311–314, 311–314, 316 pair-bonding, 102–103 Paisley Caves site, 149 Palenque site, 325, 362, 365–368, 365–368 paleoanthropology, 19, 37 See also hominin evolution; human origins paleoethnobotany, 19, 219 Paleoindian period, 148–149, 148–150, 152, 156, 157–158, 166 Paleolithic period, 32, 161, 161–162, 511 architecture of, 262 art and, 31, 79 defined, 78 divisions of, 78–80, 80, 111 fire use and, 79, 101, 161 hunter-gatherers, 79, 103, 133–134, 161–162, 211–212 technology, 78–80, 78–80 See also Homo erectus Paleozoic era, 4, 5 palisades, 284, 295 Pampa Grande site, 419 Pañamarca site, 407 panpipes, 278 pantheon, 447 Papago people, 303 paper, invention of, 485 papyrus, 462, 463 Paracas textiles, 404, 404 Paranthropus, 42, 42, 58, 66 Paranthropus robustus, 42, 44, 53 Parkington, John, 179, 180, 182 passage grave, 525, 526, 526 pastoralists, 493 patrilocal residence pattern, 354 PCR (polymerase chain reaction) technique, 116 Pearson, Michael Parker, 293, 531 pebble tools, 59, 59, 64, 79, 99, 99, 100 pectorals, 334 Penck, A., 81 Peng-tou-shan site, 236 percussion flaking, 63, 64 Pericles, 502, 541, 564 Perigordian period, 124 Perrot, Jean, 209 Petra site, 11 Petralona skull, 89 Petrie, W. M. Flinders, 234 petroglyphs, 31, 31, 143, 314, 314 petty states, 378–380 Pfeiffer, John, 134 pharaohs, Egyptian, 439, 465, 465–466, 467, 468–470, 471, 473 Philip (Macedonian ruler), 541, 557 phonetic writing, 369, 465 phosphate analysis, 10 photography, aerial, 9, 9 photomicrographs, 55, 55 physical anthropology, 19
Page I-9
phytoliths, 20 pictographs, 31, 369, 465 pigs, 53, 213, 215, 217, 223 Pima people, 303 Pincevent site, 138–139, 138–139 Pinnacle Point site, 105 pipal tree, 459 pipes, 278, 279, 279–280, 283, 298, 298 pithoi, 536, 537 pithouses, 184, 184, 185 Pizarro, Francisco, 423, 428 plant remains Abu Hureyra site, 214–216, 216 Çatalhöyük site, 229 Mehrgarh site, 233 Monte Verde site, 155–156, 254 Poverty Point site, 273 preservation of, 399 See also archaeobotany plants, 2–3 platform pipes, 279, 279 Plato, 533, 541 Pleistocene epoch (Ice Age), 4, 32 animal extinctions and, 127, 148, 148, 150, 151–152, 151–153 climate/environment, 81–83, 81–84, 89, 101, 103, 147–148, 151–152, 157 divisions of, 80 Homo erectus and, 73, 81 mammals in, 95–97, 95–97, 151–152, 151–152 Pliocene epoch, 4, 32, 42, 45 See also hominin evolution Plio-Pleistocene, 41–42, 43, 55 plows, 448, 454, 542, 555 pochteca, 382 political organization, 25 Bronze Age Europe, 534, 541 defined, 29 Egypt, 439, 462–474 Inca civilization, 391, 393–394, 418, 423–425, 429, 431–432, 433–434 Iron Age Europe, 557–558 Maya civilization, 439, 488 Mesoamerica, 361–362 native North Americans, 268, 286, 289, 318 See also ciudadelas; state development pollen samples, 17, 20–21, 186, 517 polychrome pottery, 410 polygynous mating systems, 46, 47 polymerase chain reaction (PCR) technique, 116 Pomona site, 367 Pompeii, 49 population growth birth spacing, breast-feeding and, 260–261, 264 impacts on, 479, 483, 502, 503, 542 subsistence patterns and, 264, 264 population pressure hypothesis, for agriculture origins, 206–207, 208 porphyry, 464, 465 portable art, 128, 129, 129, 131–132, 136–137, 136–137 post molds, 273 Postclassic period (Mesoamerica), 325 Postglacial Europe, 166, 169, 169–170 Postglacial period, 107, 165–168 Carrier Mills site, 167, 187–191, 187–191 climate/environment, 165, 169, 169–170 Elands Bay Cave site, 166, 179–181, 179–182 hunter-gatherers, 165–167, 179–182, 183–186 Jomon period Japan, 166, 183–186, 183–186 nutrition and, 165–167, 170, 173, 183–184, 186, 188, 196 Sannai Maruyama site, 166, 183–186, 183–186 site map, 168 Star Carr site, 170 timelines for, 167, 168
Vedbaek site, 164, 165, 166, 171–174, 171–175, 177 potassium-argon (radiopotassium) dating, 45, 48, 48–49, 50 potlatch, 316 potsherds, 234, 235 pottery, 234–235 Cardial, 507 Chinese, 475, 476 Dolni Vestonice site, 129 Egyptian, 464 Hoabinhian sites, 240 Indus civilization, 454–455 Jomon period, 166, 184, 185, 186, 186, 234, 235 Lapita, 144, 144 Mehrgarh site, 232–233 Mesoamerican, 323, 333, 335, 336, 336, 354 Mesopotamian, 442, 443, 443, 447, 448 Mexico, 234–235, 234–235 Moche civilization, 406–408, 408 native North American, 272, 272, 274, 274, 278, 286, 286, 289–290, 290, 302, 303, 303, 308, 308 Nazca civilization, 410 Neolithic Europe, 514, 514 Southwest Asia, 208, 213 Tiwanaku/Wari, 418, 418 Yangshao sites, 236–238, 238 Potts, Richard, 55 Poverty Point site, 9, 269, 272–275, 272–275, 325 Precambrian era, 4, 5 prehistory, 1, 8, 78, 268 Present Interglacial period, 4, 84, 169–170, 169–170 See also Holocene (Recent) epoch; Postglacial period Presentation Theme, 409, 409, 414 preservation, 12, 14, 399, 517 pressure flaking, 158, 158 Prezwalski horse, 96 primary context (in situ), 15 primary domestication centers, 201–202, 203, 204, 208 primates, 35 pristine (indigenous) states, 322 propaganda, 31 propagation, 199–200 provenience, of artifacts, 15 public space, 329, 387 See also architecture Pueblo Bonito site, 308–309, 309, 310 Pueblo Indians, 307 pueblos, 308 Puleston, Dennis, 363 Pulltrouser Swamp, 364 puna region, 257 punctuated equilibrium, 7 Puuc-style architecture, 374, 374–375 Pylos site, 540–541 pyramids, 269 Egypt, 439, 468–470, 468–470, 473–474 Mesoamerica, 287, 329, 332, 338, 338, 340, 474, 474 South America, 391, 395, 405–406, 405–407, 410, 411–412, 411–413 Teotihuacan site, 287, 329, 349, 350, 350, 351–353, 352–353, 355, 473, 474 Qafzeh Cave site, 121 Qin dynasty (China), 439, 483–484, 487 quarries. See extraction sites querns, 209 Quetzalcoatl (god), 373, 376 Quilter, Jeffrey, 397, 398, 399 quinoa, 400 quipu, 394, 429, 429, 451 rachis, 220, 220 radiocarbon dating, 17, 48, 121, 145–146, 145–146
Index
I-9
_
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
radiopotassium (potassium-argon) dating, 45, 48, 48–49, 50 raised-field agriculture, 363–364, 364 Rambo, A. Terry, 502 ramón trees, 363 Ramses II, (Egyptian pharaoh), 471 rank, 28 See also status differentiation Rawson, Jessica, 484 Recent epoch, 4–5 See also Holocene (Recent) epoch reciprocal exchange, 27, 28, 281–282, 383, 555 reconnaissances (unsystematic surveys), 347 record(s) archaeological, 14–15 monument, 361–362, 361–362 See also archives red ochre, 128, 129, 139, 143, 242 redistribution, 27, 281–282, 383, 474 Redman, Charles, 501 Redmond, Elsa, 345 reduction technique, 99, 100 regional settlement patterns, 24, 348, 434 regionalization, 470, 527, 542 Reindel, Markus, 410 relative dating method, 48 religion agriculture and, 207, 207 Incan, 426 Mesoamerican, 352, 384 See also Christianity; Islam; ritual/ceremony repoussé, 404 retouching, 64 rhizomes, 257 rice, 202, 236, 239, 239, 242, 263 Rice, Glen, 305 Richerson, Peter, 315 Rindos, David, 207 rites of passage, 31 ritual/ceremony art and, 134, 137 blood importance, 384, 385 China, 475, 478, 480–481 Draper site, 296–298 geoglyphs, 410 Hopewell sites, 276, 279–280 ideologies and, 31, 385 Incan, 429, 431–432 Indus civilization and, 459 Iron Age Europe, 550–551 Mesopotamia, 448 Moche Sacrifice Ceremony, 409, 409, 413, 414 Moundville site, 291 Neolithic Europe, 527 nonresidential architecture and, 329–330 potlatch, 316 sites for, 24, 134 See also human sacrifice road system, Incan, 394, 427–428, 427–428, 433 Roberts, Frank, 157 Robson, J. R. K., 244 robust skeletons, 118, 119, 193 Rock of Cashel, 562 roe deer, 170, 170 Rogers, Alan, 75 Roman cosmology, 30 Roman empire, 509, 534, 547, 552, 554, 557–558 roof comb, 359 Roosevelt, Anna, 392 Roosevelt, Theodore, 567 root crops, 200, 202, 242, 256 Ruddiman, Walter, 83
_ _
SAA. See Society for American Archaeology Sabloff, Jeremy A., 363 saches, 359, 377 Sacred Almanac, 370, 370 Sagan, Carl, 2, 574 Sahagún, Bernardino de, 373
I-10
Index
Page I-10
Sahelanthropus tchadensis, 40, 66 Sahlins, Marshall, 196, 502, 503 Sahul Strait, 142, 142 Samoa, 144 San Bartolo site, 341 San José Magote site, 325, 326–328, 326–328, 342 architecture, 327–328, 328, 329–330, 330, 342–343 human sacrifice, 327, 327, 328 San Lorenzo site, 325, 331, 331–334, 334, 335 San (Bushmen) people, 194 Sanders, William, 333–334, 350 sandstone, 272 Sannai Maruyama site, 166, 183–186, 183–186 Santayana, 563 Saqqarah site, 468, 473 sarcophagus, 366 Sassoon, Siegfried, 531 Sauer, Carl, 207, 240, 242 scalping, 292 scanning electron microscope, 220 scapulimancy, 451, 475, 476, 478 scavenging, 55 scepters, 413, 414 scheduling (resource extraction), 251 Schliemann, Heinrich, 539, 540 Schmandt-Besserat, Denise, 451–453 Schöningen site, 102, 102, 103 Scorpion, (Egyptian king), 464, 465 scrapers, 100 seal stamp, 444 season awareness, 126, 140–141, 140–141 seasonal settlements, 23 seasonality, 251 secondary domestication centers, 201 sedentism agriculture and, 207, 226, 260, 262, 263, 264 defined, 207 Postglacial period, 166, 172, 179, 196 sediment analysis, 22 seed crops, 200 Seibal site, 367 serpentine, 458 setaria grass, 253 settlement analysis, 23–24, 24, 26 settlement patterns analysis of, 24, 25, 25, 26 Mesoamerica and, 321, 347–348, 348 systematic surveys, 347–348, 348 sexual dimorphism, 40, 41, 42, 66–67 sexual division of labor, 102 sexuality, 67–68, 409 Shabik’eshchee site, 306, 308 shaduf, 471, 471 shaft graves, 538, 539 shamans, 31, 404 Shang dynasty (China), 439, 475–479, 481 Shang-shan site, 236 Shanidar site, 107, 121 shattering, 220, 246 sheep, 213, 215, 217, 226, 263 shell middens defined, 15, 181 Klasies River Mouth Caves site, 113 Ozette site, 311 Postglacial period, 167, 181, 184 shellworking, 460–461, 461 shicra, 395 Shihuangdi (Chinese emperor), 483–487 Shipman, Pat, 40, 55 Shub-ad (Sumerian queen), 449–450 sickle polish, 511 sickles, 210 side-looking airborne radar (SLAR), 364 Siemens, Alfred, 363 silk, 237 Sima de los Huesos site, 91, 91–94, 92 Simonides, 540 Simpson, George Gaylord, 36
Simpson, James, 307 Sinanthropus pekinensis, 85 Singer, Ronald, 112 Sipán site, 390, 391, 393, 411–414, 411–414 site datum point, 14 site maps Aegean Bronze Age, 532 Africa, 440, 492 Asia, 440 cave paintings, 131 China, 476 Egypt, 462 Europe, 89, 91, 510 hominin evolution, 38 Homo erectus, 77 Homo sapiens evolution, 108, 144 Indus civilization, 454 Mesoamerica, 243, 324 Mesopotamia, 441 Mississippian culture, 290 native North Americans, 271, 277, 290, 306 Neanderthals, 108 Neolithic period, 212, 224, 236 New World migration, 147 Postglacial period, 168 South America, 393, 407, 415, 419, 423, 427 Southeast Asia, 240, 488 sites analysis of, 23–24, 24 defined, 8 destruction of, 13 discovery of, 8–11 grids, 14 selection of, 12, 14 Siyaj Chan K’awaiil II (Mayan ruler), 361, 362, 362, 387 Skara Brae site, 573 skeletal evidence, 23, 192–193, 192–193 Skhul site, 121, 122 skin color, human, 75, 107 skull racks (tzompantli), 345, 345, 356 SLAR. See side-looking airborne radar slash and burn agriculture, 280, 357, 363, 364 slate, 274, 275 Smith, Bruce, 244 Smith, C. Earle, 255 Snaketown site, 269, 270, 300–305, 300–305 soapstone, 451 social hypotheses, for agriculture origins, 207, 208 social systems. See kinship systems; political organization; status differentiation Society for American Archaeology (SAA), 563, 569, 570, 571 Socrates, 541, 557 sodalities, 31 soft-hammer technique, 100 soil analysis, 10, 17–18, 22 solifluction, 128 Solis, Ruth Shady, 395 solstices, 285, 367, 426, 530–531 Solutrean period, 124 Solutrean tools, 124, 124, 125 South Africa De Hangen site, 179, 179, 180–181, 182 Elands Bay Cave site, 166, 179–181, 179–182 hominin evolution, 40–41, 41 Homo sapiens evolution, 105 hunter-gathers in, 55, 206 Klasies River Mouth Caves site, 111–113, 111–113 Swartkrans site, 52–53, 52–54 South America, 390–435 agriculture, 202, 204, 254–255, 257, 392, 394, 395, 396, 398, 407, 411, 416, 417, 418, 430–431 Amazonian, 392 Chavín de Huántar site, 392, 400–403, 400–403, 404, 434
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
Chimú civilization, 393, 419–422 climate/environment, 391, 399, 399, 416 Guitarrero Cave site, 204, 254–257, 254–257, 392 Monte Verde site, 149, 154–156, 154–156, 254 monumental architecture, 395–397, 401–403, 401–403, 434 site maps, 393, 407, 415, 419, 423, 427 state development, 322, 392, 433–434, 503 timeline, 393 Tiwanaku site, 393, 415–418, 415–418 trade, 396, 400, 417, 434 See also Inca civilization; Moche civilization South Asia. See Indus civilization Southeast Asia agriculture, 488, 491 Homo erectus, 75 Homo sapiens evolution, 142 site maps, 240, 488 state development, 439, 488–491, 488–491 Southeastern Ceremonial Complex, 289–290, 290, 293 Southern Cult, 289 Southwest Asia Abu Hureyra site, 204, 207, 212, 214–217, 215–217, 225–226, 226 agriculture, 201–202, 202, 204, 205–206, 208, 211–213, 218, 224, 224–226, 226, 234, 261, 262, 441–443 ‘Ain Mallaha site, 204, 209–210, 209–210, 262 Çatalhöyük site, 204, 227–231, 227–231 FMH, 121 Homo erectus, 103 Jericho site, 204, 207, 212, 221–223, 221–223, 227 Neanderthals, 119, 121 state development, 322, 438–439 Spanish conquest. See European-Native American contact spearpoint styles, 31 spears, 73, 102, 102 spearthrowers, 125–126, 125–126, 137 specialists, 19 speech. See languages Spence, Michael, 350, 351, 354 Spencer, Charles, 345 spindle whorls, 237 Spirit Cave site, 241 Spiro site, 283, 289 split inheritance, 420 Spy cave site, 117 Square Tower House site, 560, 560, 561 squash, 244, 247–248, 253, 258–259, 322–323 Squier, E. George, 428 stadia rods, 10 Standage, Tom, 448, 470 standardization, 456, 459 Stanish, Charles, 415 Star Carr site, 170 Starcevo site, 508 state development, 437–439 Africa, 439, 497–499 China, 439, 475–479, 482–487, 503–504 Indus civilization, 438 Mesoamerica, 321–323, 325 Mesopotamia, 438–439, 465 South America, 322, 392, 433–434, 503 Southeast Asia, 439, 488–491, 488–491 Southwest Asia, 322, 438–439 theories of, 439, 501–504 water management and, 438, 443, 465, 491, 502 See also political organization states, 30, 322 See also political organization; state development; specific states and areas status differentiation, 515 Cahokia site, 285, 286, 287 Chaco Canyon site, 310
Page I-11
chiefdoms, 315–316, 318 China, 439, 475, 477–478, 481, 482 Egypt, 436, 436, 437 grave goods and, 293–294, 294, 407, 515, 546 Hopewell sites, 269, 276–277, 279 Inca civilization, 433–434 Iron Age Europe, 548 Khmer civilization, 490 Mesoamerica, 382, 433, 434 Mesopotamia, 444, 449 Moche civilization, 407, 409 Moundville site, 291, 292, 293–294, 294 organization and, 28, 30 settlement analysis and, 23 steatite, 272, 458 Stein, Gil, 444 Steinheim skull, 89, 90 stelae, 341, 354, 361–362, 361–362, 367 stingray spines, 327 Stirling, Matthew, 331 stirrup-spout vessels, 407, 408 Stoltman, James B., 283 Stone Age, 78, 111, 516 hunting during, 112–113 organic materials from, 13, 13 time divisions for, 32, 78 See also Mesolithic period; Neolithic period; Paleolithic period stone boiling, 233 Stonehenge site, 24, 509, 525, 528–530, 528–531, 543, 555 Stoneking, Mark, 107 storage, 433, 537 food, 199, 201, 207, 222, 234, 283, 315, 407, 430–431 stratification. See status differentiation stratigraphic section, 553 stratigraphy, 15, 15–17 Çatalhöyük site, 227, 227 dating and, 15–17, 48 Sannai Maruyama site, 185, 185 Zhoukoudian site, 86, 87, 88 strip mining, 187 stucco, 339 “stuff-and-cut” eating habits, 118 styles, 31 sub-Saharan Africa agriculture, 202, 263, 263 Great Zimbabwe site, 439, 496–500, 496–500 Jenné-jeno site, 439, 493–494, 493–495 state development, 439 subsistence patterns bone chemistry and, 176–177, 176–178, 182, 253 changes in, 204 defined, 27 population growth and, 264, 264 See also agriculture; hunter-gatherers Sugiyama, Sabruro, 351 Sumer, 442–443, 449 See also Mesopotamia superposition, law of, 15 supply and demand, 281–282 surveyor’s benchmark, 14 surveys, 8–10, 9–10, 12, 33 defined, 8 systematic settlement pattern, 347–348, 348 survival of the fittest, 7 Suryavarman II (Khmer ruler), 490 Susa site, 451 Swanscombe skull, 89 Swartkrans site, 52–53, 52–54 sweat baths, 296–297 symbols fertility, 137, 137 ideologies and, 31 in medieval Europe, 29, 29 notation and, 140–141, 140–141
systematic settlement pattern surveys, 347–348, 348 Taieb, Maurice, 45 talud-tablero style, 351, 351 tampus, 428 T’ang (Chinese king), 480 Tasmania, 107, 144, 169 Tauber, Henrik, 177 Tawantinsuyu, 391, 423 See also Inca civilization taxation, redistribution and, 281, 383, 474 technology Africa, 492 Bronze Age Europe, 542, 542–543, 555 chimpanzees and, 32, 32, 63, 68 China, 487 defined, 27 hominin evolution and, 42, 55, 63–64, 63–65 Kalambo Falls site, 98 Mesopotamia, 448 Neolithic Europe, 511, 517, 523, 523 Olduvai Gorge site, 57, 57–60, 59–60, 64 Paleolithic period, 78–80, 78–80 time and, 32 Upper Paleolithic, 121, 123–126, 123–126 Zhoukoudian site, 86, 88 teeth age at death and, 192, 193, 225 hominin evolution and, 41–42, 42 nutrition and, 210, 260, 519 Tehuacán sites, 204, 243, 247–249, 250–253, 250–253 Tell as-Sawwan site, 443 Tell Brak site, 444 Tell Hamoukar site, 444 Tello, Julio, 401 tells, 214, 222–223, 223, 227, 441 temper, 240, 241 temple institution, 438, 443, 445, 449 Temple of the Great Jaguar (Tikal site), 359, 359 temples. See monumental architecture temporal markers, 234, 235 Tenochtitlán site, 356, 372, 378–380, 378–382, 382 European-Native American contact and, 321, 385 monumental architecture, 378, 378–379, 382, 382, 387 teosinte, 247–249, 247–249 Teotihuacan site, 325, 345, 349–354, 349–355 architecture, 329, 349–355, 350–353, 386 art, 354, 354–355 fall of, 372 growth of, 349–351 pyramids, 287, 329, 349, 350, 350, 351–353, 352–353, 355, 473, 474 Tepe Gawra site, 443 Terra Amata site, 101 terracotta, 439 terracotta soldiers, 439, 483–487, 483–486 test pits, 10, 15, 15 Texcoco, 380 textiles agriculture and, 212, 237 Guitarrero Cave site, 255, 256 Mesoamerican, 372, 382 native North American, 313 South American, 394, 397, 404, 404, 432 Tezcatlipoca (god), 373 Tezozomoc, Fernando Alvarado, 325 Thebes, 470, 471, 540 theories, 25, 27, 33, 106, 107, 107 Theropithecus, 98 This site, 463 tholoi, 443 tholos tomb, 540, 541 Thompson, Donald, 428, 429, 431 Thomsen, Christian, 78 Three Books of Rites, 480
Index
I-11
_
pri31057_ndx.ppI1-I12.qxd
10/15/09
10:06 AM
Three Dynasties period (China), 475–476, 482 Thucydides, 502 Tigre Complex (El Mirador site), 338, 339–340 Tikal site, 325, 339, 354, 357–362, 357–362 architecture, 329, 357–360, 357–360, 368 monument record, 361–362, 361–362 Timbuktu, 495 time, 2–5, 3, 5, 32 awareness of, 126, 140–141, 140–141 See also calendars; dating; specific eras, epochs, and periods Tiryns site, 540 tissue reconstructions, 75, 92, 106, 118, 120 Tiwanaku site, 393, 415–418, 415–418 tlachtli, 356 Tlacopan, 380 Tlaloc (Aztec deity), 378, 382 Tlatelolco site, 321, 380, 383 tobacco, 330 Tobias, Philip, 42 Toffler, Alvin, 574 Tollund Man, 14, 550–551, 551 Toltec civilization, 371–373, 371–373, 375–376 tombs. See burials; megaliths Tonga, 144 tools, 63–64, 63–65 See also technology; specific tools topographic maps, 8–9 Torralba site, 90, 97 total station, 14, 14 totem poles, 311, 312 Toth, Nicholas, 63–65 Towry, Kristy, 452 toys, 235, 386 trade Aegean Bronze Age, 532, 533–534 Africa, 439, 492, 493, 494–495, 495, 497, 499–500 Bronze Age Europe, 542–543, 543 China, 483 craft specialization and, 438, 460 defined, 27–28, 28 Egypt, 468, 470 Indus civilization, 458, 459 Iron Age Europe, 549, 557 maritime, 488 Mesoamerica, 354, 372, 382, 386 Mesopotamia, 448 native North Americans, 278, 280, 281–282, 289–290, 299, 302, 310, 316 Neolithic Europe, 512, 515, 523 South America, 396, 400, 417, 434 transhumance, 180 travertine stone, 372 Tres Zapotes site, 331–332 tribes, 30, 315 trilithons, 529, 530, 531 Triple Alliance, 380, 380 tropic level (food chain position), 177–178 tsetse fly, 493, 496 tubers, 257, 400 Tugen Hills site, 40 Tula site, 325, 371–373, 371–373, 378 tumplines, 313 Tutankhamen (Egyptian pharaoh), 467, 471, 473 Tuthmosis III, (Egyptian pharaoh), 471
_ _
I-12
Index
Page I-12
Tylor, E. B., 437 tzompantli (skull racks), 345, 345, 356 Uaxactún, 360, 361 ‘Ubaid period, 442–444, 445, 446–447 ‘Ubeidiya site, 75, 103 Uhle, Max, 405 underwater archaeology, 19, 19, 520, 521 UNESCO World Heritage Site program, 284, 564, 564, 565 Unetice culture, 543 unifacial tools, 64 unions, 29 Upper Paleolithic, 78, 79, 107, 161 season awareness, 126, 140–141, 140–141 subperiods of, 124 technology, 121, 123–126, 123–126 Upper Paleolithic art, 104, 105, 126, 161 cave paintings, 96, 127, 130–135, 130–135 portable art, 128, 129, 129, 131–132, 136–137, 136–137 Ur site, 449–450, 450, 456 uranium-series dating, 93 Uruk site, 438, 444, 446, 446–449, 449–450, 451 Ushki I site, 148 Vaenget Nord site, 174, 175 Vague Year, 370 Vallon-Pont d’Arc, 135, 135 Varna site, 508, 513–515, 513–515 vaults, 377 Vavilov, Nikolai Ivanovich, 240 Vedbaek site, 164, 165, 166, 171–174, 171–175, 177 Venus figurines, 128, 137, 137 Venus of Brassempouy, 104, 105 vertebrates, 35 Vertezöllös site, 90 vertical excavations, 14, 15–16, 15–17 Vikings, 22, 31, 509 Vinca culture, 514 Vix site, 509, 548–549, 548–549, 552 Vogelherd site, 129 wadi, 467 Wak Chan K’awaiil (Mayan ruler), 362 Walker, Alan, 41, 74 walking. See bipedalism Wallace, Alfred Russel, 6, 143 Wallace Line, 142, 142, 143 waranqa, 429 warfare China, 479, 482, 483, 484 Inca civilization, 423, 429 Iron Age Europe, 552, 554, 557 Mesopotamia, 448–450 native North Americans, 292, 299, 315 Southeast Asia, 488, 489, 491 state development and, 502, 503 Wari site, 418, 419 Warka vase, 448, 448 water management China, 483, 485 Egypt, 462–463, 465, 471, 471, 472 floodwater farming, 306, 307 Hohokam culture, 270, 302–303, 304, 304
Mesoamerica, 350, 363–364, 382 Mesopotamia, 441–442, 443, 446 South America, 407, 420 state development and, 438, 443, 465, 491, 502 Watson Brake site, 273 wattle and daub technique, 300 Waymire, Gregory, 452 weaning, 260–261 Webb, C. H., 274 Wedding Rock site, 314, 314 weights, Harappan, 456, 457, 459 were-jaguar figures, 335, 336, 336 Wessex culture, 543 wet-site excavation, 313 whaling, 311–314 Wheeler, Mortimer, 552–554 wheels, for transportation, 235, 448, 554 wheel-thrown pottery, 235, 447 White, Leslie, 69 White, Tim, 40, 42 Wilcox, David, 301, 305 Willey, Gordon, 347 Williamsburg, Virginia, 571 Wing, Elizabeth, 257 witches, 31 Wittfogel, Karl, 350, 502 Wolf, Eric R., 372 woodhenges, 285, 529, 531 Woolley, Leonard, 449–450 woolly mammoths, 72, 73, 95, 95–96, 127, 127–129, 129, 148, 148, 149, 152 World Heritage Site program, UNESCO, 284, 564, 564, 565 Wright, Henry T., 503 writing, 8 Aegean Bronze Age, 533, 533, 534 China, 451, 453, 475, 478–479, 484 Egypt, 451, 453 Indus civilization, 458, 459 Maya civilization, 325, 341, 369–370, 369–370, 386 Mesopotamia, 448, 451–452, 451–453 See also hieroglyphs; pictographs Wymer, John, 112 Xia dynasty (China), 475–476 Xianyang, 439, 479, 482–487 Yangshao sites, 236–238, 236–238 Yax Nuun Ayiin I (Mayan ruler), 361, 362 Yaxchilán site, 361, 362 Yik’in Chan K’awaiil (Mayan ruler), 359 yin and yang, 480 Ying Zheng (Chinese king), 483 Zapotec civilization, 328, 330, 346, 369 Zeder, Melinda, 225 Zeribar site, 215 Zhengzhou, 476 Zhou dynasty (China), 475, 479, 480–481, 482–483 Zhoukoudian site, 85–88, 85–88, 103 ziggurats, 223, 445, 446, 447 Zinjanthropus, 58, 62 zoomorphic designs, 237, 238, 404 Zoser (Egyptian pharaoh), 468, 473
Th eDiTion
Th
eDiTion
offers a beautifully illustrated, full-color, site-byThe sixth edition of site survey of prehistory. Through five successful editions, instructors and students alike have been captivated by the book’s rich visual presentation, research, interpretations, and theories from the field. The new edition has been revised in accordance with both new discoveries in archaeology and the interests of readers.
gary Feinman
T. Douglas Price Features of the Sixth Edition The newest discoveries, research, and interpretations in archaeology have been added to the sixth edition. The visual program of the book has been further improved upon with new photos and line drawings throughout.
The first half of the book includes: new information on the relationship between archaeology and anthropology and historical archaeology (Ch. 1); a new section on argon-argon dating and new information on Atapuerca and the first Europeans (Ch. 2); a new hypothesis on meteors and Pleistocene extinctions (Ch. 3); new dates and a discussion of early, fully modern humans from Pinnacle Point, SA and a new discussion of first Americans (Ch. 4); new material on nitrogen isotopes in relation to the study of the paleo diet (Ch. 5); and new information on the domestication of plants and animals from Peru, Near East, China, Africa, and Europe (Ch. 6).
Price Feinman
ISBN 978-0-07-353105-2 MHID 0-07-353105-7 9 0 0 0 0
9
780073 531052 www.mhhe.com
MD DALIM 1049758 09/25/09 CYAN MAG YELO BLACK
New material in the second half of the book includes: new findings concerning the introduction of corn into the American Southwest (Ch. 7); early Mayan writing and new perspectives on the competition between different Mayan states during the Classic period (Ch. 8); new information on the mapping and size of Angkor (Ch. 10); the latest information on the Iceman, including the use of strontium isotopes to determine his place of origin, and exciting new data from the Stonehenge Riverside project, including interpretation of life and death contexts (Ch. 11). Finally, Chapter 12 has been expanded with additional information on career choices.